Recent advances in CO2 capture and reduction

Kecheng Wei a, Huanqin Guan a, Qiang Luo b, Jie He *bc and Shouheng Sun *a
aDepartment of Chemistry, Brown University, Providence, Rhode Island 02912, USA. E-mail: ssun@brown.edu
bDepartment of Chemistry, University of Connecticut, Storrs, Connecticut 06269, USA
cPolymer Program, Institute of Materials Science, University of Connecticut, Storrs, Connecticut 06269, USA. E-mail: jie.he@uconn.edu

Received 25th May 2022 , Accepted 16th July 2022

First published on 18th July 2022


Abstract

Given the continuous and excessive CO2 emission into the atmosphere from anthropomorphic activities, there is now a growing demand for negative carbon emission technologies, which requires efficient capture and conversion of CO2 to value-added chemicals. This review highlights recent advances in CO2 capture and conversion chemistry and processes. It first summarizes various adsorbent materials that have been developed for CO2 capture, including hydroxide-, amine-, and metal organic framework-based adsorbents. It then reviews recent efforts devoted to two types of CO2 conversion reaction: thermochemical CO2 hydrogenation and electrochemical CO2 reduction. While thermal hydrogenation reactions are often accomplished in the presence of H2, electrochemical reactions are realized by direct use of electricity that can be renewably generated from solar and wind power. The key to the success of these reactions is to develop efficient catalysts and to rationally engineer the catalyst–electrolyte interfaces. The review further covers recent studies in integrating CO2 capture and conversion processes so that energy efficiency for the overall CO2 capture and conversion can be optimized. Lastly, the review briefs some new approaches and future directions of coupling direct air capture and CO2 conversion technologies as solutions to negative carbon emission and energy sustainability.


Kecheng Wei received his B.S. degree (2017) in Chemistry from the University of Science and Technology, China. He joined Prof. Shouheng Sun's group in 2017 and is currently a Ph.D. candidate in the Chemistry Department of Brown University. His research is in the shape-controlled synthesis of nanocatalysts for fuel cell reactions and CO2 reduction.

Huanqin Guan obtained his B.S. degree from Peking University in 2018. He then joined Prof. Shouheng Sun's group and is currently a Ph.D. candidate in the Chemistry Department at Brown University. His research interests involve nanocatalyst synthesis and their applications in green chemistry reactions, and CO2 capture and conversion.

Qiang Luo received his B.S. degree in Material Science and Engineering from Shaanxi Normal University in 2019. Currently, he is a Ph.D. student at the Department of Chemistry at the University of Connecticut under the supervision of Prof. Jie He. His research interests include the synthesis and applications of mesoporous materials and CO2 reduction.

Jie He earned his B.S. and M.S. degrees in Polymer Materials Science and Engineering from Sichuan University and his Ph.D. in Chemistry from the Université de Sherbrooke in 2010. After working with Professor Zhihong Nie as a postdoctoral fellow at the University of Maryland, he joined the faculty of the University of Connecticut where he is currently an Associate Professor of Chemistry. His group focuses on the design of hybrid materials of polymers and inorganic materials (metal ions, clusters, and nanoparticles) being capable of catalyzing the activation of H2O, O2 and CO2 as inspired by nature.

Shouheng Sun received his Ph.D. in Chemistry from Brown University in 1996. He joined the IBM T. J. Watson Research Center (Yorktown Heights, New York) first as a postdoctoral fellow (1996–1998) and then as a research staff member (1998–2004). In 2005, he returned to Brown University as a tenured Associate Professor and was promoted to full Professor in 2007. He is now the Vernon K. Krieble Professor of Chemistry and Professor of Engineering. He served as Associate Editor of the Royal Society of Chemistry journals Nanoscale/Nanoscale Advances (2012–2021) and is a Fellow of the Royal Society of Chemistry. His main research interests are in chemical synthesis and self-assembly of nanoparticles for catalytic, magnetic, and biomedical applications.


1. Carbon cycle and CO2 emission

Carbon is the chemical backbone of life on Earth. Carbon compounds regulate the Earth's temperature, make up the food that sustains us, and provide the energy that drives the global economy. The carbon cycle in nature is the global flow of carbon through the atmosphere, oceans, terrestrial biosphere, and lithosphere in various forms, such as carbon dioxide, organisms, limestone, coal and oil, as shown in Fig. 1A.1 Two main cycles are the land–atmosphere cycle and the ocean–atmosphere cycle.2,3 The land–atmosphere cycle occurs through two main drivers: photosynthesis and respiration. In the photosynthesis process, CO2 is absorbed from the atmosphere and converted into fuels by plants or microbes, while in the respiration process CO2 is produced as the final product from biological activities. In comparison with the land–atmosphere cycle, the ocean–atmosphere cycle plays a vital role in carbon storage because the ocean contains 50 times more carbon than the atmosphere.4,5 The driving mechanism of the ocean–atmosphere cycle is the difference in the partial pressure of CO2 between the ocean and the atmosphere. This pressure varies with ocean temperature and local marine photosynthesis. The lower the ocean temperature, the smaller the carbon emissions.6 In all, nature balances these cycles well in equilibrium, maintaining the healthy evolution of life. Over the past century, however, fossil fuels have been massively consumed for energy uses (Fig. 1B).7,8 This has resulted in a dramatic increase in atmospheric CO2, and as a result, caused a series of environmental issues, including global warming, acid rain, ocean acidification and rising sea levels (Fig. 1C).7–10
image file: d2nr02894h-f1.tif
Fig. 1 (A) The schematic highlights carbon fluxes through inland waters and includes pre-industrial and anthropogenic fluxes. Values are net fluxes between pools (black) or rates of change within pools (red); units are Pg C per year; negative signs indicate a sink from the atmosphere. Gross fluxes from the atmosphere to land and oceans, and the natural (Nat) and anthropogenic (Ant) components of net primary production—the net uptake of carbon by photosynthetic organisms—are shown for land and oceans. Gross primary production (GPP) and ecosystem respiration (R). (B) Energy source in the past and forecast from 1970 to 2050, and (C) CO2 concentration in atmosphere, global temperature, sea level. Adapted from ref. 1 and 9 with permission. Copyright 2009 Nature Publishing Group and 2012 Chelsea Green Publishing.

2. CO2 capture

To combat anthropogenic CO2 emission and to make our lifestyles sustainable, we must develop neutral or even negative carbon emission technologies. One such technology is CO2 capture and storage.11,12 Recent studies have shown that the key barrier that limits the broad use of this process is the high energy penalty associated with CO2 capture,12 which is aggravated by the fact that about half of the annual CO2 emission is generated from widespread industrial sites. In 1999, Lackner introduced the concept of direct air capture of CO2 to mitigate climate change, and it is now broadly defined as direct extraction of CO2 from ambient air.13–16

A key step to successful CO2 capture is to develop efficient adsorbents to bind CO2 from air. An ideal CO2 adsorbent should have high selectivity and adsorption capacity for CO2, low heat of adsorption (Qst), high recyclability, good thermal and chemical stability, fast kinetics and high cost-effectiveness (Fig. 2).17 The energetics of the CO2 capture process is about the chemical bonding nature between CO2 and an adsorbent, which can involve both weak physisorption and strong chemisorption. Such binding strength is defined by the isosteric heat of adsorption Qst (kJ mol−1). For a given adsorbent, a high Qst value indicates an energy-intensive CO2 regeneration process once it is captured, whereas a low Qst value may compromise the CO2 adsorption capacity. Furthermore, a good adsorbent should have high selectivity for adsorbing CO2 from a mixture of gases, especially from air, and have high thermal, chemical, and water stability to achieve high CO2 capture efficiency under different operational conditions. The rate of CO2 uptake should also be kinetically fast, the capture process should be easily engineered to large scale, and the overall cost for the capture process should be economically practical. Here we summarize some representative adsorbents that have been studied extensively for CO2 capture, including aqueous hydroxide, solid alkali carbonates, organic amines, and porous materials.18,19


image file: d2nr02894h-f2.tif
Fig. 2 Principal criteria for designing an ideal sorbent for CO2 capture. Reprinted from ref. 17 with permission. Copyright 2020 Royal Society of Chemistry.

2.1. Hydroxide-based adsorbents

Due to the relatively low concentration of CO2 (412 ppm) in the atmosphere, CO2 capture is usually carried out using chemical adsorbents with a strong CO2 binding affinity. A common adsorbent is calcium hydroxide solution, which can react with CO2 and form calcium carbonate as precipitate. The calcium carbonate can then be separated and dried for storage. The captured CO2 can be accessed through a process known as calcination – the decomposition of calcium carbonate to form calcium oxide with CO2 being released as a concentrated stream. Calcium hydroxide is then regenerated in a slaking process via hydration of calcium oxide, forming a recyclable loop.14 Many different types of adsorption devices, from traditional stagnant pools, packed towers to modern spray towers, have been designed and developed.20 Alternatively, solid inorganic bases are used for ultra-dilute CO2 removal. Fig. 3 plots the equilibrium partial pressure of CO2, pCO2,eq, as a function of temperature (T), for various single-metal oxide sorbents. Combinations of T and pCO2 above the respective pCO2,eq curves imply the material exists as carbonate, whereas below the curve, the material's thermodynamically stable state is its oxide form.21 The thermodynamic properties of the CaO–CaCO3 system enable the direct capture of CO2 from ambient air (pCO2 = 4 × 10−4 bar). To release CO2 from CaCO3, high temperatures (>900 °C) are generally required to obtain a pure stream of CO2 (pCO2 ≈ 1 bar). Despite the convenient chemistry involved in the process, dealing with a large volume of air, hydroxide solution, and metal carbonate decomposition can impose heavy energy cost due to the high temperature required to regenerate the metal oxide adsorbents and to release CO2.
image file: d2nr02894h-f3.tif
Fig. 3 Equilibrium partial pressure of CO2, pCO2,eq, as a function of temperature for alkali (green) and alkaline earth (blue) metal oxide–carbonate systems. Horizontal gray lines indicate pCO2 of 0.05, 0.15, and 1 bar, respectively. Reprinted from ref. 21 with permission. Copyright 2008 and 2021 American Chemical Society.

2.2. Amine-based adsorbents

Amines are another common adsorbent employed for CO2 capture. Using aqueous solutions of amines to capture CO2 has been extended to commercial uses to remove CO2 from CO2-rich natural gas streams.22 To date, amine adsorbents employed for direct air capture have been supported on solids to improve amine stability and recyclability. The strength of chemisorption between an amine and CO2 ensures selective CO2 uptake even at low CO2 partial pressures, which makes the solid-supported amine adsorbents highly suitable for the direct air capture of CO2.

In a dry condition, CO2 reacts with either a primary amine (eqn (1)) or secondary amine (eqn (2)) to produce an ammonium carbamate. When moisture is present, the reaction yields ammonium carbonate or bicarbonate (depending on the pH) (eqn (3) and (4)).14 Amine-containing sorbents have been divided into three classes: class 1 amine adsorbents are prepared by impregnating amines into the pores of a support; class 2 amine adsorbents are formed by covalently bonding amines to the walls of porous materials via silane linkage; and class 3 amine adsorbents are derived from polymerization of amines in situ to form polyamine structures tethered to the inner walls of the porous support.14,23–25Fig. 4 shows some representative examples of these amine adsorbents. After adsorption, CO2 can be released from the adsorbent by an inert gas flow to drive the reaction equilibrium towards gaseous CO2.14,26 An advantage of the amine adsorbent over the metal oxide one is its selective adsorption of CO2 over H2O, making it useful in the humid environment.27

 
CO2 + 2RNH2 ⇄ RNH3+ + RNHCOO(1)
 
CO2 + 2R1R2NH ⇄ R1R2NH2+ + R1R2NCOO(2)
 
CO2 + R1R2NH + H2O ⇄ R1R2NH2+HCO3 ⇄ R1R2NH2+CO32− + H+(3)
 
CO2+ R1R2R3N + H2O ⇄ R1R2R3NH+HCO3 ⇄ R1R2R3NH+CO32− + H+(4)


image file: d2nr02894h-f4.tif
Fig. 4 (A) Molecular structures of commonly used amines for class 1 and class 2 direct air capture sorbents. (B) Schematic representation of the three main routes used for functionalization of porous supports with amine moieties. Reprinted from ref. 14 and 26 with permission. Copyright 2016 American Chemical Society and 2016 Georgia Institute of Technology.

2.3. Adsorbents based on metal organic frameworks (MOFs)

The cleavage of adsorbed CO2 in the formation of carbonate and carbamate is endothermic; and it requires a large energy input to regenerate in case of a strong adsorbent. Weak physisorption of CO2 by porous materials has been explored extensively as an alternative to chemisorption to improve the energy efficiency of the capture process. Significant research progress has been made in CO2 capture by micro and mesoporous materials, including metal–organic frameworks (MOFs), zeolites, zeolitic imidazole frameworks (ZIF), and porous polymers.13,17,28,29 MOFs consist of three-dimensional coordination polymer networks, constructed by the combination of metal ions/clusters with organic linkers/ligands. Fig. 5 shows crystalline structures of some well-known MOFs.30 Their CO2 adsorption power can be tuned more conveniently by specific surface areas, pore volumes, pore sizes, metal centres, and surface functional groups, which make them especially attractive for selective CO2 capture from a stream of gas mixture.
image file: d2nr02894h-f5.tif
Fig. 5 Crystalline structures of some well-known MOFs. Reprinted from ref. 30 with permission. Copyright 2019 Elsevier.
2.3.1. CO2 adsorption via metal-binding in MOFs. As the pore structures of MOFs are sensitive to the adsorption of CO2, functionalization of the inner and outer surface of the MOFs can tune their adsorption power. A typical approach is to make MOFs with unsaturated open metal sites (UOMSs), which can be prepared by heating or vacuuming of the solvent-coordinated metal cations in MOFs.30,31 These exposed metal-coordination sites can build an electric field around them, providing the desired driving force for CO2 adsorption. M–CO2 binding is realized by direct interaction between the antibonding dz2 orbital and the lone electron pairs on the oxygen in the CO2 molecule. For an early transition metal cation with d electrons less than 4, its antibonding d-orbitals tend to bind to CO2 more strongly. But for a late transition metal cation, its antibonding orbitals can be filled up, weakening its binding with CO2.32 For example, Mg2(dobdc) (H4dobdc = 2,5-dihydroxyterephthalic acid), Mg-MOF-74, and CPO-27-Mg structures with open metal sites bind to CO2 at a fixed angle in a uniaxial fashion, as shown in Fig. 6A.33 Their CO2 binding energies are around 67.2 kJ mol−1 (Fig. 6B).34 In the presence of the early transition metal (Ti or V) cation, their binding energies increase to 73.2–80.2 kJ mol−1, while in the presence of the late transition metal (Cr or Zn) cation, their binding energies drop to 32.2–50.8 kJ mol−1 (Fig. 6B).
image file: d2nr02894h-f6.tif
Fig. 6 (A) Schematic illustration of CO2 uniaxial rotation at the open Mg2+ site in Mg2(dobdc). Gray – C, red – O, and green – Mg atoms; H atoms are omitted for clarity. The blue circle is the arbitrary rotation axis. (B) CO2 adsorption energy plotted over Q/r computed for the metal–CO2 oxygen distance and the tetrazole nitrogen–CO2 carbon distance. The vertical line is to guide the eye. Reprinted from ref. 33 and 34 with permission. Copyright 2012 and 2014 American Chemical Society.
2.3.2. Enhancing CO2 adsorption via functionalization of MOFs. MOFs modified with functional groups can change their surface properties and CO2 adsorption power. A common strategy to modify MOFs is to add polar or amine-based moieties to the structure to enhance their interactions with CO2, especially under low-pressure conditions. However, this enhancement needs to be regulated very carefully as the strong interaction with CO2 also makes it difficult to regenerate the MOF adsorbent. Therefore, this functionalization should enable MOFs to show high CO2 adsorption affinity, capacity, and selectivity, but low Qst.

As an example, isoreticular MOF (IRMOFs)-74-III was functionalized with a series of organic linkers –CH3, –NH2, –CH2NHBoc (Boc: tert-butyloxycarbonyl), –CH2NMeBoc, –CH2NH2, and –CH2NHMe via the Suzuki–Miyaura coupling reaction (Fig. 7A).35,36 All the modified MOFs, except the ones containing the protective Boc groups, showed high and similar CO2 adsorption behaviours at 25 °C/800 Torr, as shown in Fig. 7B.36 However, at low pressure, primary amine- and secondary amine-functionalized MOFs (IRMOF-74-III-CH2NH2 and IRMOF-74-III-CH2NHMe, respectively) outperformed the other modified MOFs (Fig. 7C).36 A second CO2 isotherm after evacuation of the sample at room temperature for 2 h and a third cycle with a heat treatment at 120 °C under vacuum (10 mTorr) for 1 h was recorded (Fig. 7D and E).36 The decrease in CO2 uptake on the second cycle and recovery upon heat treatment indicated the presence of strongly bound CO2. Further exploration of IRMOF-74-III-CH2NH2 using dynamic CO2 adsorption under dry (16% CO2; 84% dry N2) and wet (16% CO2; 84% wet N2) conditions showed a negligible difference in the uptake rates for the CO2 adsorption, suggesting the unique structural selectivity towards CO2. 13C NMR studies showed that the capture was realized by chemisorption between CO2 and the functionalized organic linkers, forming carbamate ions and carbamic acids for IRMOF-74-III-CH2NH2 and IRMOF-74-IIICH2NHMe, respectively. Incorporation of diamine groups into the same MOF to form IRMOF-74-III-(CH2NH2)2 could provide an even higher CO2 adsorption power at 25 °C/800 Torr (75 cm3 g−1) than that of IRMOF-74-III-CH2NH2 (67 cm3 g−1).37


image file: d2nr02894h-f7.tif
Fig. 7 (A) Synthetic pathway for the functionalized organic linkers used in the synthesis of IRMOF-74-III, in the preparation of –CH3 (5a), –NH2, (5b), –CH2NHBoc (5c), and –CH2NMeBoc (5d) functionalized linkers. On the right is shown a schematic representation of the IRMOF-74-III pore as functionalized with the organic linkers 5a–5d and post-synthetic deprotection of Boc groups. Color code: C in gray, O in red, functional groups in purple, Mg as blue polyhedra. (B) Comparison of CO2 uptake at 25 °C for IRMOF-74-III-CH3 (gray), –NH2 (green), CH2NH2 (red), –CH2NHMe (blue), –CH2NHBoc (purple), and –CH2NMeBoc (cyan). (C) Expansion of the low-pressure range (>1 Torr). Carbon dioxide isotherms at 25 °C for IRMOF-74-III-CH2NH2 (D) and –CH2NHMe (E). Uptakes for samples after activation (first cycle), after first CO2 uptake (second cycle), and after 120 °C heating for 1 h for regeneration (third cycle) are shown in circles, triangles, and squares, respectively. Reprinted from ref. 36 with permission. Copyright 2014 American Chemical Society.
2.3.3. MOF pore size-dependent CO2 adsorption. The pore size is a third common parameter that can be applied to control MOF's CO2 adsorption capability and selectivity. It is possible to synthesize MOFs with microporosity, mesopores or macropores, which can be controlled by the nature of metal precursors and organic linkers used during synthesis.38,39 For example, MOFs with pore size of 2.6, 2.4, and 2.2 nm could be synthesized using cobalt–organic linkers with slightly different configurations (IR-MOF-74-III).40 The benzene rings were termed as pore size tuners and the CO2 adsorption of the three MOFs was enhanced as the pore size decreased from 2.6 to 2.2 nm. The competing adsorption of water could be suppressed by narrowing down the pore size as suggested by computational calculations and experimental demonstration on MOF-74 by inserting 2,4,6-tri(4-pyridyl)-1,3,5-triazine (tpt) into its hexagonal channels.41

3. CO2 reduction

Despite the fact that CO2 capture is important to solve CO2 emission problems, to realize energy sustainability, CO2 must be converted back into chemical fuels, which requires the controlled reduction and protonation of CO2. This process is unfortunately energetically uphill due to the high activation energy needed to break the stable O[double bond, length as m-dash]C bonds and the apparent difference in free energy between CO2 and the final products. For this reaction to be economically viable, suitable catalysts with high catalytic activity, selectivity and stability must first be developed to achieve energy-efficient reduction of CO2. Many chemistry processes, including thermochemistry, electrochemistry, photochemistry, and biochemistry processes, have been studied for CO2 reduction. In this section we highlight the recent advances in thermo- and electro-catalytic reduction of CO2.

3.1. CO2 activation

CO2 is a very stable molecule, with a bond dissociation energy of 525.9 kJ mol−1 and ionization potential of 13.777 eV, making CO2 activation difficult and costly.42 One-electron reduction of CO2 is believed to be the first step to initiate the reduction and other reaction processes that convert CO2 to reusable forms of carbon. The electronic structure of CO2 in different charge states can be summarized in its Walsh diagram (Fig. 8A).43 In the ground state of neutral CO2, the highest occupied molecular orbital (HOMO) is the fully occupied 1πg orbital. An excess electron will be accommodated in the 2πu orbital, which is stabilized by bending the molecule, leading to a deviation of the molecular symmetry from D∞h to C2v. The singly occupied molecular orbital (SOMO) in this radical anion is of a1 symmetry, with an OCO angle calculated to be 138°, and it can be described as pseudo-antibonding. At the same time, the bonding 1πg orbital transforms into a2 and b2 orbitals that have been characterized as largely nonbonding.44 Consistent with the pseudo antibonding nature of the HOMO of CO2, its C–O bond length (124 pm) is greater than that of the neutral CO2 (117 pm).
image file: d2nr02894h-f8.tif
Fig. 8 (A) Walsh diagram of CO2, with illustrations of the highest occupied molecular orbitals of the anion (top) and the neutral (bottom). (B) Structural motifs of metal–CO2 interactions. Reprinted from ref. 43 and 44 with permission. Copyright 2014 Abingdon: Taylor & Francis and 2018 Annual Reviews.

The free CO2 radical anion is metastable and has been observed in mass spectrometry with measured lifetime up to milliseconds.44 The radical anion can be stabilized by interaction with a matrix or by solvation. The solvated CO2 radical anion has been observed in bulk solutions as well as in (CO2)n (n = 6–13) cluster ions.44 While the first electronic excited state of neutral CO2 is in the deep ultraviolet (UV), the radical anion has its lowest excited state in the near UV range. The electronic absorption band of CO2 is at about 235 nm and CO2 can dissociate upon excitation and lose its excess electron by charge transfer, making it challenging to fully characterize CO2.

Understanding the binding between CO2 and a metal surface is of great importance for developing a metal catalyst to catalyse CO2 reduction reaction. CO2 can bind to metal atoms via different binding motifs, as summarized in Fig. 8B.44 These modes are abbreviated as η1-C, η1-O, η2-(C,O), and η2-(O,O), where superscripts denote the number of bonds between the metal atoms and bound CO2, and the chemical element symbols describe the atoms directly interacting with the metal. Electron reduction of metal–CO2 leads to the formation metal–CO2 cluster anions, [M(CO2)n] that can serve as simplified models for studying CO2 binding to metal atoms present on catalyst surfaces.45–47

3.2. Thermal reduction of CO2

The reduction of CO2 in a thermo-catalytic process has attracted much attention as it not only reduces CO2 emission, but also directly produces value-added chemicals and fuels.48 To transform CO2 to downstream products, its thermodynamic reaction barrier must be overcome.49 Using H2 as a high-energy reactant to reduce CO2 has been a common approach, as H2 can be generated from water electrolysis by renewable (solar or wind) electricity.48 Therefore, this conversion of CO2 by catalytic thermo-hydrogenation is one of the most attractive approaches to sustainable energy and a carbon-neutral cycle (Fig. 9A).50
image file: d2nr02894h-f9.tif
Fig. 9 (A) Conversion of CO2 to chemicals and fuels through hydrogenation. (B) Schematic illustration of cycles between RWGS, CO2-FT and methanol synthesis. Reprinted from ref. 50 and 51 with permission. Copyright 2016, 2018 Royal Society of Chemistry.
3.2.1. Reversible water–gas shift (RWGS) reaction to CO. CO is considered as the most crucial intermediate in CO2 conversion as it can be coupled in methanol synthesis and Fischer–Tropsch (FT) synthesis of various chemicals and fuels (Fig. 9B).51 CO is generally produced by the reversible water–gas shift (RWGS) reaction, in which CO2 is hydrogenated under a high-temperature and high-pressure condition. However, this reaction quickly reaches its equilibrium, and as a result, the reaction has a low conversion yield (23%) at 300 °C and 1 MPa.52 Two mechanisms have been reported to explain the CO2 hydrogenation to CO. The first one is a redox mechanism, which is usually observed on the surface of Cu-based catalysts. CO2 is reduced by Cu0 to form CO*, which is desorbed from the surface to form CO product, and Cu+ is then reduced back to Cu0 by H2 with water being formed as a byproduct.53 This is further supported by density functional theory (DFT) calculations and Fourier-transform infrared spectroscopy (FTIR) spectroscopy studies over a Cu/ZnO catalyst.54 The CO2 hydrogenation may also follow the formate pathway, in which CO2 is first converted to formate that is further dehydrated to form CO.55

Metals on oxide supports are considered as promising catalysts as the metal centres could easily dissociate H2, which is followed by transfer of H* to CO2 adsorbed on the oxide support.56 Various catalysts based on transition metals on different oxide supports have been studied for the RWGS reaction. Among them, Cu or Pt-based catalysts supported on CeO2 are the most extensively studied.57 In studying monometallic and bimetallic Pt-based catalysts on different oxide supports for selective CO2 conversion to CO, it was found that active metal controlled the product selectivity, while the support effect dominated the activity of CO2 conversion.58 For the monometallic Pt catalysts, a reducible support (CeO2) showed higher activity than an irreducible support (γ-Al2O3) because of the increased oxygen vacancies found in the CeO2 structure, which are beneficial for oxygen exchange with CO2. Among the bimetallic Pt-based catalysts supported on CeO2, PtCo showed the highest CO selectivity with little CH4 production due to the weak binding of CO on the metal surface (Fig. 10A).58 Based on the d band theory, the CO/CH4 ratio selectivity increases when the values of the d-band centre move towards more negative values for the Pt, Co, and Ni-based catalysts on either CeO2 or γ-Al2O3 supports (Fig. 10B).58 Such correlation between CO selectivity and metal d-band centre is potentially helpful for predicting selective CO2 reduction catalysts.


image file: d2nr02894h-f10.tif
Fig. 10 (A) Pt–Co/CeO2 for CO2 hydrogenation to CO. (B) Effect of d-band centre on ratio of CO to CH4 production at 10% conversion. For ease of comparison, open and solid symbols represent catalysts with and without Ni, respectively. Reprinted from ref. 58 with permission. Copyright 2013 Elsevier.

In addition to Pt, other precious metals, such as Ir, Ru, Rh and Pd, are reported to be highly active hydrogenation catalysts.59 Alternatively, Cu, Fe and Ni-based catalysts are also being explored for large-scale RWGS.60 Cu/CeO2 was found to be especially active as a RWGS catalyst at low temperature (300 °C) and ambient pressure, reaching 100% CO selectivity.61 The enhanced activity was attributed to synergies of Cu nanoparticle (NP) and CeOx support in their redox behaviours and oxygen vacancies (Fig. 11A).61In situ Ce L3-edge XANES measurement for CeO2 and Cu–CeO2 supported on mesoporous silica SBA-15 (denoted as SCe and SCuCe, respectively) was performed during sample reduction at 300 °C in H2 (Fig. 11B and C).61 No change was observed in SCe after reduction, while there was partial reduction of Ce4+ to Ce3+ for SCuCe. Similarly, Cu structure change upon CO2 treatment at 120 °C was detected by in situ X-ray absorption spectroscopy (XAS) measurement (Fig. 11D and E).61 The spectra corresponded well with the Cu2O standard, suggesting that Cu0 species in SCuCe-re are oxidized to Cu+ species via CO2 treatment at 120 °C.61 The rapid desorption of CO from Cu+–CO intermediate at the reaction temperature led to product formation, which was followed by facile reduction of Cu and Ce by hydrogen spillover. The results suggest that the synergistic effect between oxygen vacancies and Cu redox property is essential for the oxide-supported Cu catalyst to show high RWGS activity and selectivity.


image file: d2nr02894h-f11.tif
Fig. 11 (A) Schematic illustration of synergy of Cu/CeOx for CO2 hydrogenation. (B–E) XAS spectra of Cu and Ce oxidation state change during reaction (note that SCe = SBA-supported CeO2 and SCuCe = SBA-supported Cu–CeO2). Adapted from ref. 61 with permission. Copyright 2018 American Chemical Society.
3.2.2. Thermal reduction of CO2 to methanol. Methanol is an important chemical feedstock for uses in combustion engines, fuel cells, and in the synthesis of downstream value-added products, such as dimethyl ether and hydrocarbons.62 The so called “methanol economy” is an indispensable and promising component in the carbon capture and conversion process to achieve a carbon neutral cycle.63 In fact, 140 million tons of methanol were produced in 2018 and its production is expected to double by 2030.64 Conventionally, methanol is produced from syngas (CO + H2) over a Cu/ZnO/Al2O3 catalyst at 200–300 °C and 3.0–5.0 MPa, but further studies indicate that CO2-blended syngas shows higher reaction rates than syngas alone under the same reaction conditions.62,65 As a result, direct hydrogenation of CO2 to methanol has been a hot trend of research. Oxide-supported Cu catalysts are popular choices for methanol synthesis from CO2. In the Cu/ZnO catalyst system, the high catalytic activity is attributed to the special Cu/ZnO interfacial and CuZn surface alloy effects, as confirmed by studying CuZn(111) and ZnO/Cu(111) catalysis.66 In this study, CuZn was found to undergo surface oxidation under reaction conditions and the surface Zn was transformed into ZnO. The catalysis showed a volcano-plot trend between methanol production and ZnO coverage on Cu(111) (Fig. 12A and B).66 Similarly, CuZn(211) catalysis was further enhanced once the CuZn surface was partially covered with ZnO. The catalysis enhancement was attributed to the strong metal–support interaction, which strengthens the surface binding to intermediates and increases the catalytic activity.67
image file: d2nr02894h-f12.tif
Fig. 12 (A, B) CO2 conversion to methanol. (C) Initial TOFs of methanol formation over Cu⊂UiO-66 and Cu on UiO-66. The reaction rates were measured after 1 h. Reaction conditions: 7 sccm of CO2, 21 sccm of H2, 10 bar, and 175 °C. Adapted from ref. 66 and 70 with permission. Copyright 2017 American Association for the Advancement of Science and 2016 American Chemical Society.

Since CO2 conversion to methanol is sensitive to catalyst structure, it is important to maintain the catalyst dispersion and prevent the catalyst from sintering and deactivation under the reaction conditions.68 Various strategies have been proposed to solve the deactivation issues, including the use of reducible supports and encapsulation of Cu in metal organic frameworks (MOFs).69,70 For example, a Cu-MOF-based composite catalyst was prepared by encapsulating Cu within the Zr-based UiO-66 porous structure.70 The stabilized Cu showed much higher activity toward methanol formation (Fig. 12C).70 Additionally, the SiO2-supported Ni–Ga intermetallic catalyst was found to be more active than the conventional Cu/ZnO/Al2O3 catalyst for the CO2 reduction to methanol at ambient pressure.71,72 A specific stoichiometric ratio (Ni5Ga3) was required in the catalyst formulation, which was stabilized by SiO2, to achieve high selectivity. Interestingly, redox-active In2O3 was also found to be a promising catalyst component with high methanol selectivity and remarkable stability due to its ability to form oxygen vacancies and metallic In in the reaction process.73 Once the In2O3 catalyst was supported on ZrO2, its catalytic activity was further improved and the methanol selectivity reached 99.8% with a CO2 conversion of 5.2% and long-term stability of 1000 h under the industrially relevant reaction conditions.

3.2.3. Fischer–Tropsch reaction. The C2+ hydrocarbons, such as alkanes, olefins and liquid fuels, are important for today's chemical and energy industries. For example, olefins are currently produced on the order of 200 million tons per year and widely used in synthetic rubbers, plastics and cosmetics.74 However, these hydrocarbons are traditionally generated from non-renewable fossil fuels, which results in large amounts of CO2 emission.75 Ideally, CO2 can be used as a precursor for the synthesis of these hydrocarbons.76

The FT reaction is a common route for the transformation of syngas (CO + H2) to C2+ hydrocarbons. To achieve the direct hydrogenation of CO2, two successive reaction steps need to be incorporated into one reaction system: the reduction of CO2 to CO via RWGS reaction and hydrogenation of CO to hydrocarbons via FT reaction.77 In the two-step reaction process, the CO conversion (up to 87%) is much higher than the CO2 conversion (up to 45%). Therefore, improving the catalytic efficiency of the CO2 conversion has been an important target.77–80 Fe-based catalysts have been widely used in CO2 hydrogenation because of their high activity for both RWGS and FT synthesis.81 Fe catalysts with alkali metal promoters are reported to significantly enhance the selectivity towards long-chain hydrocarbons.82 These alkali metals, especially K, promote Fe catalysis by weakening the affinity with H2 and enhancing the adsorption of CO2 and CO intermediate.83 Different promotional effects were observed by combining a Fe-based MOF catalyst with various elements (Fig. 13A).83 K was found to enhance the olefin selectivity drastically from 0.7% to 36% (Fig. 13B).83 CO2 and H2 chemisorption measurement showed that CO2 uptake was enhanced while H2 adsorption was weakened upon K addition, leading to stronger Fe–C interaction and higher selectivity toward olefins. The obtained C2–C4 olefin space time yield (STY) of the Fe/C + K (0.75) catalyst was among the best catalysts published (Fig. 13C).83


image file: d2nr02894h-f13.tif
Fig. 13 (A) Synthetic strategy for the Fe-based catalysts by carbonization at 600 °C in N2 and wetness impregnation (W. I.). (B) Effect of different promoters on CO2 hydrogenation performance. (C) C2–C4 olefin space time yields (STY, mmol gcat−1 h−1) obtained for the Fe/C + K (0.75) catalyst at 350 °C compared with the best catalysts available for CO2 hydrogenation. Adapted from ref. 83 with permission. Copyright 2018 American Chemical Society.

In addition to alkali metals, transition metal components, such as Cu and Co components, were also found to promote Fe-catalysed CO2 hydrogenation to hydrocarbons.84,85 Cu is known to be a highly active catalyst for methanol synthesis from CO2, but when it combines with Fe, it enhances Fe catalysis for both RWGS and CO hydrogenation by suppressing CH4 formation and promoting C2–C7 production.84 The catalyst support is also an important factor to increase the selectivity for light olefins. For example, the ZrO2-supported K–Fe (K–Fe/ZrO2) catalyst exhibited much higher selectivity to lower olefins than the SiO2-supported one;86 the carbon-coated Fe-catalyst was much better dispersed and stabilized, and was highly active for the CO2 conversion at atmospheric pressure with higher selectivity to C2–C4 olefins.87 In addition, methanol has also been studied as a starting precursor for synthesis of olefins. It too requires two reaction steps: CO2 hydrogenation to methanol and methanol conversion to hydrocarbons as described in recent reviews.88,89

Despite the great promise demonstrated from thermal reduction of CO2 to value-added chemicals, these thermal reactions do require the use of high temperature and high pressure, which makes it challenging to stabilize the catalysts in the reaction conditions and to lower energy consumption.

3.3. Electrochemical reduction of CO2

Electrochemical CO2 reduction reaction (CO2RR) is an appealing alternative to thermal reduction for converting CO2 to value-added chemicals as the reaction can be promoted by renewable electricity under ambient conditions, and be catalysed more selectively by catalyst engineering, as illustrated in Fig. 14A.90 The electrochemical CO2RR on the surface of a metal catalyst is generally divided into three steps: CO2 adsorption, charge transfer, and product dissociation. Each of these three steps plays an important role in controlling catalyst selectivity and final product distribution.91 The CO2RR pathways have been studied extensively to understand various products detected from CO2RR. Fig. 14B is just an example to show these complicated pathways leading to the formation of C1 and C2 products.91 The commonly accepted key reaction steps are CO2 binding, protonation and reduction to *COOH, which can be further hydrogenated to form formate, or dehydrated to form CO that can either be released from the catalyst surface or function as a key intermediate for the next steps of hydrogenation and C–C coupling to C1 and C2+ products. The mechanism leading to the formation of C1 product is relatively simple. In contrast, the processes leading to C2+ products are much more complicated. Recent studies have focused on capturing and identifying the reaction intermediates, such as *COCO, *CHCHO, *COCOH, that produce C2+ products.92
image file: d2nr02894h-f14.tif
Fig. 14 (A) Schematic illustration of sustainable energy cycling based on electrochemical CO2RR. (B) Proposed pathways to C1 and C2 + products from electrochemical CO2RR. Adapted from ref. 90 and 91 with permission. Copyright 2020 American Association for the Advancement of Science, 2021 American Chemical Society.
3.3.1. Metal nanoparticle catalysis. Metal nanoparticles with large surface areas and controlled surface structures have been studied extensively as catalysts for CO2RR.93–99Fig. 15 summarizes some representative nanoparticle catalysts that are selective for CO2RR to CO.93 Ultrathin Au nanowires of about 2 nm in width and hundreds of nm in length were found to be among the most active and selective catalyst for the CO2 reduction to CO.99 The CO selectivity is sensitively dependent on the length of the nanowires. The 500 nm Au nanowires showed the onset potential of CO2 reduction to CO at −0.2 V (with 37% FE) but reached 94% FE and mass activity (1.84 A gAu−1) at −0.35 V. DFT calculation revealed that both COOH and CO preferentially bind to the edge site on the Au nanowires, with COOH binding marginally (0.04 eV) stronger than that on the Au(211) edge but CO binding 0.23 eV weaker than that on the Au13 corner, suggesting that nanowire surface with maximal edge sites facilitates CO2 reduction to COOH and further to CO.99 In addition to Au, Ag, Pd, SnO2-coated Cu, and Ni–N were also found to be selective in catalysing CO2RR to CO, as summarized in Fig. 15.100–103 When Pd, In, Sn, and Bi nanoparticles were employed as catalysts for the CO2RR, formate (HCOO) was the main product.104–107
image file: d2nr02894h-f15.tif
Fig. 15 Summary of some representative nanocatalysts based on Au (A), Ag (B), Pd (C), Sn (D) and Ni (E) for electrochemical CO2RR to CO. All potentials are vs. RHE. NP denotes nanoparticle and FE is the reported faradaic efficiency. Adapted from ref. 93 with permission. Copyright 2019 Cell Press.

Compared with the formation of CO and formate, selective reduction of CO2 to C2 products has been challenging, and Cu has been the major component that is required to catalyze the C–C formation.92,95 Recent studies have suggested that the key active components are Cu–Cu2O mixtures, as observed in the CO2RR studies on partially oxidized Cu electrode.108–111 Cu(I) and residual subsurface oxygen species are considered to play important roles towards enhanced performance. The oxidation state of Cu can be reversibly transformed between Cu(0) and Cu(I) under the electrochemical reaction conditions. The presence of Cu(I) and Cu(0) significantly improves the kinetic and thermodynamic processes of CO2 activation and *CO dimerization. In situ spectroscopy studies, such as electrochemical liquid transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS) and XAS studies, have shown that the catalyst surface undergoes dynamic structural changes under CO2RR conditions. For example, over the electroreduction time period, CuO nanosheets were seen fragmenting into smaller species and floating in the liquid layer (Fig. 16A).112In situ grazing incidence X-ray absorption spectroscopy (GIXAS) and X-ray diffraction (GIXRD) were also used to study thin Cu electrode (50 nm thick) and to characterize the near-surface structure of the electrode under the CO2RR conditions (Fig. 16B and C).88,113 It was found that during the catalytic reaction, the surface of the polycrystalline Cu electrode was partially oxidized to Cu2O. The co-existence of Cu(0) and Cu(I) on the catalyst surface during the CO2RR was further proved by operando time resolved XAS.114 It is now commonly believed that there is a synergistic effect between Cu(0) and Cu(I), which promotes the C–C coupling of intermediates in the reaction process, favouring the formation of C2+ products. The role played by the Cu(I) effect in enhancing CO2RR to hydrocarbons was further supported by the Cu3N nanocube-catalysed CO2RR for the formation of C2H4 as a major product.115 This high selectivity to C2H4 was attributed to the Cu(I) stabilization by N and Lewis basicity of N on the Cu(100) facet, facilitating C–C coupling and C[double bond, length as m-dash]O/C–O hydrogenation.


image file: d2nr02894h-f16.tif
Fig. 16 (A) Schematic overview (timeline) of the experimentally observed evolution of the CuO nanosheet morphology probed by the in situ TEM E-chip flow cell, H-cell, and flow cell electrolyser. (B) XANES at the Cu K-edge of the CuO nanosheet catalyst after CO2 reduction for different minutes. (C) Grazing incidence X-ray diffraction of Cu(pc) at a probe depth of 2.6 nm before and after releasing the applied potential. Adapted from ref. 112 and 113 with permission. Copyright 2020 American Chemical Society and 2021 Nature Publishing Group.
3.3.2. Single-atom catalysis. Single-atom catalysts, with isolated metal atoms dispersed on conductive carriers, have demonstrated excellent catalytic performance in many chemical reactions. These catalysts integrate the benefits of both homogeneous and heterogeneous catalysts, and provide an ideal platform for optimizing chemical reactions via their easily controllable coordination sites and electronic structures, strong metal–support interactions, as well as their maximal atom utilization. The electrochemical reduction of CO2 over single metal atom sites can be traced back to the 1970s when cobalt and nickel phthalocyanines were first found to be active for CO2 reduction.116 Since then, metal–organic complexes with well-defined M–Nx sites have been extensively studied for CO2RR with high catalytic performance and durability.117 In these M–Nx structures, both C and N coordinated to M also show important synergy effects (via electronic polarization) on the M catalysis to facilitate CO2 activation and further reactions. For example, C–ZnxNiy ZIF-8 catalysts with undercoordinated Ni–Nx sites (x < 3) showed much enhanced CO2-to-CO activity and selectivity compared with the Ni–phthalocyanine one with well-defined Ni–N4 sites.118 DFT calculations revealed that the free energy for *COOH formation was lower on the Ni–Nx sites than on the Ni–N4 sites. This low coordination effect on catalysis enhancement was also observed from the Co–N catalysts.119 When the Co–N coordination number was decreased from 4 to 2 (Fig. 17A and B), the Co–N2 sites showed the best CO2RR performance with both high activity and selectivity towards the formation of CO (Fig. 17C and D).119 More and more single-atom catalysts are emerging to show promising catalysis for the CO2 reduction to C-products beyond CO. These results have been nicely summarized in several recent reviews.117,120,121
image file: d2nr02894h-f17.tif
Fig. 17 (A) XAS spectra confirm the atomic dispersion of Co atoms in Co–N2, Co–N3, and Co–N4, and suggest the lowest N coordination number in Co–N2. (B) XPS of all four samples. (C) (a) Linear sweep voltammetry (LSV) of Co–N2, Co–N3, Co–N4, and Co NPs and pure carbon paper as background. (D) CO faradaic efficiencies at different applied potentials. Adapted from ref. 119 with permission. Copyright 2018 Wiley-VCH.

Up to now, various catalysts have been demonstrated to show promising catalysis for CO2RR to value-added chemicals under ambient conditions. Their catalysis performance is better understood at the atomic scale, and many factors, including atomic composition, atom oxidation states and coordination environment, have been identified as important to improving catalysis activity and selectivity. Despite these advances, controlling catalytic CO2RR to a C2+ product is still a challenging task due to the strong competition from the reaction pathways that lead to the formation of C1 products. Key factors to maximize C2+/C1 ratios need to be identified, and catalyst structures developed for such catalysis should also stay stable in the CO2RR condition.

3.4. Catalyst–electrolyte interface engineering for CO2 reduction

As electrochemical reactions occur at interfaces of catalysts and electrolytes, and as the CO2RR needs the presence of protons to form hydrocarbons, interface engineering to control proton concentration and hydrophobicity is equally important for fast binding and reduction of CO2. Recent advances in surface chemistry and spectroscopy also offer new opportunities to probe reaction mechanisms of CO2RR at the interfaces and, in turn, guide the design of such interfaces for catalysis enhancement.122–125 Electrochemical systems that can be used to optimize gas transport,126–129 electrolyte functions,130–133 intermediate detection,134–139 and reaction pathway engineering,140–142 have been rapidly developed to improve the overall catalytic performance. In this section, we highlight how interfacial engineering can be applied to optimize electrolyte and catalyst surface ligand effects to enhance CO2RR catalysis.
3.4.1. pH gradient, cation, and anion effects. Most electrocatalytic CO2RRs are operated in an electrochemical reaction system with aqueous electrolyte. The electric double layer formed between the electrolyte and the electrode contains key intermediate species that can dominate the mass transport and catalytic evolution process of CO2.143 A number of parameters can be chosen to optimize the electric double layer structure and to control the local environment on the catalyst surfaces, including electrolyte buffer capacity,144 anion/cation types and concentrations,145 localized pH,146 and proton/CO2 accessibility.147,148 Therefore, the selection of an appropriate electrolyte is of critical importance to tune the CO2RR activity and selectivity. For example, when operating in an aqueous electrolyte, the reduction of protons, commonly known as the hydrogen evolution reaction (HER), competes with CO2RR; and it, in turn, lowers the overall electrocatalytic efficiency towards CO2RR.149 However, because of the continuous consumption of protons through proton-coupled CO2RR and HER, the accumulation of OH near the surface of a catalyst results in a higher local pH, creating a pH gradient between the interfacial area of the catalyst and electrolyte.150 The presence of this pH gradient can affect mass transport of different reaction species, such as OH, CO2, HCO3 and CO32−, and as a result, dominate the reactions pathways.144,146,151

The CO2RR selectivity can be improved by increasing the CO2 concentrations and by inhibiting the HER near the electrode–electrolyte interfaces. Due to the relative low solubility of CO2 in aqueous electrolyte, a high local pH is required to increase the local CO2 concentration and to suppress HER.152 This was better demonstrated when a mesoporous Au-inverse opal (Au–IO) structure served as the catalyst for CO2RR. In the CO2RR condition, the partial current density related to CO2RR was increased with reduction potentials regardless of the thickness of the Au–IO film (Fig. 18A), while that related to HER was decreased initially before increase only slightly at more negative potentials, and, more importantly, the thicker the Au–IO film, the smaller the partial current density (Fig. 18B).153 Here a beneficial pH gradient was created in the pores of the Au–IO film, which enhanced CO2 adsorption and conversion to CO, but limited proton diffusion and HER (Fig. 18C).153 This was further confirmed by in situ electrochemical analysis, showing the pH changes near the electrode surfaces.150,154 Another way of promoting CO2RR is to increase the CO2 pressure, and therefore the CO2 concentration, as demonstrated in the Cu-catalysed CO2RR to ethylene (FE 44%) when the CO2 pressure was set at 9 atm during the electrolysis.155 This improved selectivity to ethylene was attributed to the increase in *CO concentration in the initial catalysis steps, promoting *CO–*CO coupling and hydrogenation.


image file: d2nr02894h-f18.tif
Fig. 18 Specific activity for CO (A) and H2 (B) with different thickness of Au–IO: 0.5 μm (green triangles), 1.6 μm (blue circles), and 2.7 μm (red squares). The samples were evaluated in CO2-saturated 0.1 M KHCO3 electrolyte, pH 6.7. Error bars represent standard deviations of three independently synthesized Au–IO samples for each thickness. (C) Scheme represents the mesostructure-induced pH gradient during CO2RR. Adapted from ref. 153 with permission. Copyright 2015 American Chemical Society.

Metal cations would accumulate near the surface of electrodes under reductive potentials, forming an electric double layer, which could affect the mass transport of CO2. As shown in Fig. 19A, the constructed electrode–electrolyte interface is assigned to the inner Helmholtz plane (IHP) within which intermediate species are populated, and outer Helmholtz plane (OHP) beyond which the hydrated cations are present.143 Under a reductive potential, the hydrated cations are attracted to the electrode surface due to the Coulomb attraction, participating in the chemical reactions, modifying the catalyst surface electronic structure, and even blocking active sites on the catalyst surface. Therefore, the catalytic performance of a catalyst on CO2RR can be highly dependent on the electrolyte. For example, when a Ag electrode was studied for CO2RR to CO, it was found that CO FEs were only around 40% when the Li+- and Na+-based electrolyte was used as the reaction medium, but the FEs reached 85–90% when the electrolyte contained a larger cation, such as K+, Rb+ or Sc+ (Fig. 19B).156 Compared with smaller cations, which are strongly hydrated, the larger cations are weakly hydrated and more accessible to the surface of the electrode, leading to the decrease in their pKa and increase in the localized CO2 concentration. A similar cation size effect was also observed when a Cu electrode was studied for CO2RR in 0.1 M MHCO3 electrolyte. Increasing the cation size from Li+ to Cs+ in the electrolyte, the FE for H2 was decreased, but the FE for C2H4 and C2H5OH was increased (Fig. 19C).157 It was believed that the larger cations helped to stabilize the polar species, such as *CO2, *CO, and *OCCO, more efficiently in the reduction condition, favouring their further coupling and hydrogenation (Fig. 19D). As a comparison, CH4 FE was rarely affected by the cation sizes due to the negligible cation interaction with the nonpolar *H and *CHO intermediate species.


image file: d2nr02894h-f19.tif
Fig. 19 (A) Simplified schematic illustration of the electric double layer composed of the inner Helmholtz plane (IHP) and outer Helmholtz plane (OHP) with chemical equilibria involved. (B) Faradaic efficiencies (FEs) for CO and H2 produced over Ag at −1 V vs. RHE in CO2-saturated 0.1 M MHCO3 (M = Li, Na, K, Rb, Cs) electrolyte. (C) Faradaic efficiencies (FEs) for C2H5OH, C2H4, CH4, and H2 produced over Cu at −1 V vs. RHE in CO2-saturated 0.1 M MHCO3 (M = Li, Na, K, Rb, Cs) electrolyte. (D) Schematic illustration of the local electric field created by cation at the catalyst interface and stabilized OCCO intermediate. Adapted from ref. 143, 156 and 157 with permission. Copyright 2020 Royal Society of Chemistry and 2016, 2017 American Chemical Society.

Similarly, anions in electrolytes can also affect the CO2RR performance of metal catalysts. These anions, for example halides, can function as soft bases to bind to Au and Cu strongly to modify the catalyst surface structure or morphology during the CO2RR, as demonstrated in the CO2RR catalysis of plasma-activated Cu foil.158 It was found that I ions enhanced the reactivity dramatically (lowered the onset potential) as compared with Br and Cl ions, and the total FE for C2–C3 products (ethylene, ethanol, and propanol) reached 65% at −1.0 V (vs. RHE). I ions were thought to be strongly adsorbed on the electrode surface, enhancing the CO2 binding through the formation of I–C bonds. Anions can also regulate the pH change near the catalyst surface, affecting the catalyst's CO2RR performance.159

3.4.2. Surface ligand effects. Adding ligands on the surface of catalysts offers a powerful way to control the interface of catalyst–electrolyte.160 The inspiration is from nature where the catalytic efficiency of metalloenzymes heavily relies on the coordination environment of metal sites, e.g., protein frameworks in both first and second coordination spheres. Protein frameworks, despite not being catalytically active by themselves, are an essential component in tuning the activity and selectivity of metal sites. Modifying metal catalysts with surface ligands, therefore, can also enhance electrocatalytic performance toward CO2RR. Such modification is usually achieved by covalent or non-covalent binding of organic surfactants to metal surfaces.161–164 Surface ligands can boost the intrinsic catalytic activity of metal catalysts by reducing CO2 activation barriers,165–167 by changing mass transport during CO2RR,168–170 and/or by defining the local environment to suppress byproduct formation (e.g., HER).171

The common organic ligand used for metal surface modification is thiol in the form of R–SH, where R represents an organic substituent. –SH has strong bonding affinity with all catalytically active metal surfaces. –S is a softer base than –O and can bind to a Group 10 or 11 metal even more strongly to impact its catalysis for CO2RR.172 There have been numerous studies in modifying catalyst surfaces with thiols to improve the CO2RR selectivity.168,173,174 One example is to modify a polycrystalline Au film electrode with three different thiols, 2-mercaptopropionic acid (MPA), 4-pyridylethylmercaptan (4-PEM), and cystemine (CYS), and to study the thiol effects on the Au catalysis for CO2RR.175 Such modifications did not improve Au catalysis for CO2RR to CO (the surface coverage generally reduced the Au catalysis selectivity to CO), but they changed other parts of the Au catalysis: the 4-PEM-modified Au showed improved selectivity to formate (from 10% FE on Au to 22% on PEM-Au), while the MPA-modified Au showed nearly 100% FE towards H2 and the CYS-modified Au was more active (not more selective) for generating CO and H2. These Au catalysis changes upon the surface modifications were attributed to the proton-induced desorption mechanism associated with pKa of the thiol ligands, as illustrated in the 4-PEM-modified Au catalysis for the improved selectivity to formate (Fig. 20A), in which 1e reduction of pyridine to pyridinium also helped to bind and reduce CO2 to facilitate the second proton binding to CO2 and its conversion to formate.176 Not surprisingly, MPA with the smallest pKa promotes HER.


image file: d2nr02894h-f20.tif
Fig. 20 Comparison of partial current density and FE for thiolate ligands on polycrystalline Au: (A) proposed reaction mechanism of the formate production at 4-PEM and Au interfaces. (B) Schematic of Au electrodes with 1-methylimidazolium-terminated SAMs (IL-2, IL-6, IL-8, and IL-12). (C) FE of C2 products on both wettable and hydrophobic Cu dendrites at the total current density of 30 mA cm−2. Adapted from ref. 175, 171 and 168 with permission. Copyright 2017 American Chemical Society, 2015 Royal Society of Chemistry and 2019 Nature Publishing Group.

Another example is to use the thiol-terminated imidazolium to improve Au catalysis for the formation of ethylene glycol (CH2OH)2.171 When the Au electrode was modified with different imidazolium–SH ligands (Fig. 20B), the Au catalysis showed the ligand length-dependent CO2RR catalysis selectivity with 1-(-2-mercaptoethyl)-3-methylimidazolium bromide (IL-2)-modified Au exhibiting highest FE (87%) towards ethylene glycol.171 Such enhancement in selectivity to ethylene glycol was attributed to more efficient coupling of imidazolium aldehyde intermediates in the reaction condition. In the presence of a longer ligand chain on the Au surface, the interaction between imidazolium and Au gets weaker, limiting the charge transfer for the formation imidazolium aldehyde intermediates.

Surface ligand modification can also be used to control the microenvironment of catalytic sites and to impact catalysis efficiency. When modifying the catalyst surface with a hydrophobic ligand, the surface area becomes hydrophobic, which allows CO2 to accumulate, creating a triphasic interface of gas–electrode–electrolyte.168 For example, when modified with 1-octadecanethiol (ODT), the Cu dendritic surface became superhydrophobic with a water contact angle of 153°. Such a hydrophobic dendrite entrapped more CO2 near the Cu surface, more efficiently improving the Cu catalysis of CO2RR to C2 products (Fig. 20C).168

Amine ligands have also been broadly used not only to stabilize metal NPs in their synthesis but also to modify metal surfaces for catalysis improvement.98,99,177–181 The presence of amine groups at metal surfaces provides numerous Lewis base centers that can further improve CO2 adsorption near these metal surfaces via the amine–CO2 interaction.182 The amine ligand effect was well demonstrated in a comparative study of Ag catalysis for CO2RR once a Ag electrode was modified separately with oleylamine (OLA), oleic acid (OA) and 1-dodecanethiol (DDT). The OLA-modified Ag was found to show the highest selectivity to CO (FE 94.2%) across a broad range of potentials, while the OA- and DDT-modified Ag demonstrated only 89.1% and 71.0% FECO, respectively (Fig. 21A).183 In studying the amine ligand effect on Au catalysis for CO2RR, Au NPs supported on graphene oxide (rGO) were grafted with propylamine (PA), hexylamine (HA), OLA, ethylenediamine (EDA) or polyethyleneimine (PEI), respectively.182 It was found that amines with a linear structure favored the CO2RR to CO, and the longer the chain, the higher the CO FE (Fig. 21B).182 As a comparison, bulky branched amines can block the catalyst's active sites and prevent CO2 from interacting with the metal surface, lowering the CO2RR selectivity.


image file: d2nr02894h-f21.tif
Fig. 21 (A) FECO of OLA-, OA- and DDT-modified Ag NPs supported on carbon black (Ag/C). (B) FECO (column) and CO current density (circle) of different Au catalysts at −0.7 V (vs. RHE). Adapted from ref. 183 and 182 with permission. Copyright 2017 American Chemical Society and 2018 Wiley-VCH.

Despite the evident effect of these thiols/amines on metal catalysis, the long-standing catalyst stability issue in the CO2RR condition remains. To stabilize the NP catalysts more efficiently in the CO2RR condition, N-heterocyclic carbene (NHC) ligand has been introduced.162,164,184,185 NHCs bind with metals through the lone electron pair on C to form a strong C–metal σ bond,186–188 which has been applied to modify the surfaces of a variety of metals.189,190 More importantly, the σ-donation of NHCs enriches the charge density on metal surfaces, further promoting metal binding with electrophile CO2.186 For example, Au NPs modified with sterically bulky 1,3-bis(2,4,6-trimethylphenyl)imidazol-2-ylidene (Cb) (Fig. 22A) showed a much higher FECO (83%) and current density (7.6-fold) than the plain Au NPs (FECO = 53%) at the potential of −0.57 V.164 The tridentate NHC ligand timtmbMe (Fig. 22B)-modified Pd electrode showed not only high selectivity for CO2RR but also stability.191 In the presence of timtmbMe, the Pd catalyst exhibited a larger total current density and positively shifted onset potentials relative to the parent palladium foil (Fig. 22C).191 The onset potential for CO2RR appeared at −0.12 V, which is about 265 mV, positively shifted relative to that of the unmodified Pd. The FE of C1 products increased from the initial 23% to 86% (with timtmbMe, 82% of formate and 4% of CO, Fig. 22D) at −0.57 V vs. RHE with a 32-fold increase in current density.191 The tridentate NHC-modified Pd also showed much improved stability as evidenced from the steady product FE in the 6 h electrolysis period.


image file: d2nr02894h-f22.tif
Fig. 22 (A) Surface modification of Au NPs with 1,3-bis(2,4,6-trimethylphenyl)imidazol-2-ylidene (Cb) through ligand exchange. (B) Schematic illustration for the tripodal NHC functionalization of Pd surfaces. (C) CV scans of Pd and Pd-timtmbMe electrodes in CO2-saturated 0.5 m KHCO3 at pH 7.3. (D) FE of C1 products (CO for unmodified Pd; HCOO- and CO for tripodal NHC-modified Pd) on unmodified Pd electrodes and tripodal NHC-modified Pd electrodes at different potentials. Adapted from ref. 164 and 191 with permission. Copyright 2016 American Chemical Society and 2018 John Wiley and Sons Ltd.

Metal catalysts modified with polymer NHCs have shown significant catalytic enhancement in CO2RR. While small-molecule ligands can vary the surface properties, polymer ligands form a protective coating layer of 10–50 nm that could “gate” the accessibility of catalytic metal NPs. Monodentate and multidentate polymer NHC ligands were first studied to stabilize metal catalysts under reductive potentials and to improve the CO2RR selectivity.162 The multidentate polymer NHC ligand poly(vinylbenzyl N-methylbenzyl N-heterocyclic carbene) (PVBMB-NHC57, P1) was synthesized using quaternization of N-methyl benzimidazole with poly(vinylbenzyl chloride) (PVBC). The monodentate NHC-terminated polystyrene (PS65-NHC, P2) was prepared from the end-group functionalization of the halogen-terminated one through atom transfer radical polymerization (ATRP). After counterion exchange with KHCO3, the two imidazolium-ended polymers could graft to Au NPs (∼14 nm) at relative high grafting density, 1.3 and 0.9 chains per nm2 for P1 and P2, respectively (Fig. 23A).162 When catalysing the CO2RR, the Au-P1/C and Au-P2/C showed both higher activity and selectivity than the Au/C due to their more efficient role in suppressing HER (Fig. 23B).162 The polymer-modified Au NPs also demonstrated much improved stability, as shown in the change of the electrochemical active surface area (ECSA) of Au NPs during a 2 h electrolysis at −0.9 V (Fig. 23C).162 The citrate-capped Au NPs showed only 24.7% ECSA retention after the 2 h electrocatalysis, while the polymer NHC-modified Au NPs had ∼75% ECSA retention. Even after 11 h electrolysis, the Au-P1/C still had a FECO of 86% while the unmodified Au NPs only had <10% FECO left. The polymeric NHC-binding strategy could be applied to Pd/C (Fig. 23D), which showed the desired enhancement in both selectivity (FECO was increased from 45% to 60%) and stability (ECSA retention was improved from 10% to 91% after 2 h electrolysis).162 As a control, Pd/C modified with thiol-terminated PS ligands and other ligands (Fig. 23D and E)162 were all less stable than the Pd/C modified with polymer NHC.


image file: d2nr02894h-f23.tif
Fig. 23 (A) Schematic illustration of synthesis of P1, P2 and surface modification of NPs (yellow). (B) LSV curves measured in 0.1 m KHCO3 at a scan rate of 10 mV s−1 for all three samples. (C) ECSA retention at −0.9 V for different electrolysis times of NHC-modified Au NPs. (D) ECSA of Pd catalysts before and after CO2 reduction at −1.26 V for 2 h with various ligands as shown in (E). Adapted from ref. 162 with permission. Copyright 2019 John Wiley and Sons Ltd.

Very recently, a nanoparticle/ordered-ligand interlayer (NOLI) was proposed and applied to enhance CO2RR efficiency.169 The NOLI structure was created by the collective dissociation of bound ligands (alkylphosphonate) from a dense assembly of metal (Au, Ag or Cu) NPs. Under the reductive potentials, covalently bonded ligands detached but were maintained on the surface through the non-covalent interactions between ligands in the densely packed assembly, as illustrated in Fig. 24A.169 Consequently, this allows K+ to transport onto the catalyst surface to balance the overall charge, creating a pseudocapacitive pocket interlayer. Specifically, the ligand chains form a hydrophobic domain around the pocket that facilitates the diffusion of CO2, while inhibiting the diffusion of water/protons, into the NOLI structure, favouring CO2 conversion over the HER.192 The Ag–NOLI improved the activity and selectivity towards CO formation dramatically in CO2RR, whereas the turnover and selectivity drop to a level similar to Ag foil when the ligand layer was removed (Fig. 24B and C), supporting the importance of the NOLI structure in the selective CO2-to-CO transformation.169 This NOLI structure was demonstrated to be highly active and selective across several metals with up to 99% CO selectivity and onset overpotentials as low as 27 mV. Interestingly, even without strong chemical binding, the ligand density (with respect to the NP surface area) remains relatively stable throughout electrolysis (Fig. 24D).169 DFT calculations reveal that the specific configuration for the NOLI facilitates the bending of the adsorbed CO2 molecule, thus promotes the rate-limiting step of the polarization of non-polar CO2 with an electron transfer to form the intermediate *CO2˙.


image file: d2nr02894h-f24.tif
Fig. 24 (A) Formation of a NOLI and a metal–NOLI catalyst for selective electrocatalysis. Blue chains on the metal NPs represent chemically bonded alkylphosphonic ligands. Upon applying a negative bias on the assembled NPs, the ligands collectively dissociate from the metal surface during NP fusion and transit to a reversible physisorption state (explicitly shown by the emphasized yellow phosphonate head group). Vpos and Vneg indicate a positive (anodic) and a negative (cathodic) polarization of the metal particles, respectively. The ligand layer maintains its stability through the non-covalent interactions of the alkyl tails (blue) in an ordered configuration (indicated by the purple double-headed arrows). The resultant metal–NOLI catalyst provides a unique catalytic pocket for selective CO2 electro-conversion (C, black; O, red). (B) CO selectivity and (C) specific current density of Ag-NOLI, Ag foil and Ag particles after the NOLI is removed from Ag-NOLI, at −0.68 V vs. RHE. (D) Ligand density of Ag–NOLI estimated from XPS throughout CO2 electrolysis. Adapted from ref. 169 with permission. Copyright 2021 Nature Publishing Group.

4. Coupled CO2 capture and conversion

Both CO2 capture and CO2 conversion processes are considered as promising strategies to reduce CO2 emissions, therefore mitigating global warming and other associated environmental concerns. However, most of the present CO2 reduction studies, either thermal or electrochemical conversion, are based on pure CO2 as the feedstock, and there exist large gaps between the capture and conversion processes. In a conventional CO2 capture and conversion process, CO2 is first captured from either ambient air or flue gas by various capture technologies. Then CO2 is desorbed, compressed and utilized in the preparation of value-added products by chemical reduction reactions.193 From the perspective of the whole system, however, the desorption and compression steps are energy-intensive, imposing a large energy penalty on the processes of CO2 capture and conversion.194 Therefore the combination of CO2 capture and conversion has been suggested in a single integrated CO2 capture and utilization process. The energetics comparison between independent and coupled CO2 capture and conversion processes is shown in Fig. 25.195 Dilute CO2 can be captured through the formation of CO2–X adduct for both processes; however, the independent one requires an additional regeneration step to produce pure CO2 for electrocatalysis, causing an extra capture “overpotential” energy. If both pathways have the same energetic level for CO2RR intermediates, integrated configurations could achieve lower overall energy requirements due to the energy saving through bypassing the capture media regeneration step. Therefore, integrating the capture and conversion processes is critical to decrease the cost and make the overall process energy efficient.
image file: d2nr02894h-f25.tif
Fig. 25 CO2 capture and conversion energetics for type-I and II (red) and type-III (green). Adapted from ref. 195 with permission. Copyright 2021 Nature Publishing Group.

To date, coupling between CO2 capture and conversion via thermo- and electro-catalysis has been studied only in a few reports. Therefore, this section summarizes the recent progress made in these two catalysis areas. The integrated capture and conversion were first demonstrated in 2013.196 In the report, polyamines and amidine bases were used for CO2 capture in alcohol solvents, and the capture products were subsequently hydrogenated to obtain alkylammonium formate salts by a Ru-based homogeneous catalyst at 40 bar H2. The best conversion performance was achieved when CO2 was captured by 1,5-diaza bicycle [4.3.0] non-5-ene and glycol to form alkyl carbonate, which was then reduced to formate with 55% yield. It should be noted that the captured CO2 can facilitate hydrogenation and yield better performance in comparison with equivalent free gaseous CO2, indicating the CO2 activation upon capture with amines. CO2 could also be captured by amines in aqueous media and subsequently converted to alkylammonium formate salts (Fig. 26A).197 The major advancement of this capture/conversion system over the previously reported one is the use of a biphasic solvent, shown in Fig. 26B. CO2 can be captured as carbamate or bicarbonate in aqueous amine solution, while the catalyst is dissolved in an organic solvent. This allows easy separation of the catalyst and formate product, and higher reaction rate due to good solubility of the captured CO2 in water. The captured CO2 was selectively converted to formate (up to 95% yield) in the presence of homogeneous Ru- and Fe-based pincer complexes.


image file: d2nr02894h-f26.tif
Fig. 26 (A) CO2 capture and conversion to HCOOH. (B) Catalyst recycling by phase separation. Adapted from ref. 197 with permission. Copyright 2016 Royal Society of Chemistry.

Recently, the coupled capture and conversion further led to the synthesis of methanol at a 79% yield.198 In this process, CO2 was captured by a short-chain polyamine, pentaethylenehexamine (PEHA), to form ammonium carbamate and bicarbonate, which was further hydrogenated at 155 °C and 50 bar of H2 for 55 h in the presence of a pincer Ru-complex catalyst (Fig. 27A). It should be noted that formate and formamide are essential intermediates for amine-assisted hydrogenation of CO2 to CH3OH. Similarly, alcohol-assisted CO2 hydrogenation to methanol via formate ester has also been studied extensively, and was further extended to a new approach of CO2 capture and conversion to methanol via alkali–metal hydroxides in ethylene glycol (Fig. 27B).199 Different from amines, hydroxides do not suffer from volatility and oxidative degradation issues. More importantly, due to their high CO2 affinity, these hydroxides have high efficiency for direct air capture of CO2. In the one-pot system, CO2 from atmospheric air was efficiently captured by an ethylene glycol solution of KOH to form alkyl carbonate intermediate, which was hydrogenated at 140 °C and 70 bar of H2 for 72 h to form to methanol at a 100% yield. Such a high yield synthesis of methanol was attributed to the facile hydrogenation of the ester intermediate. Also in the process, hydroxide was partially re-generated and could be used for the next round of CO2 capture and conversion.


image file: d2nr02894h-f27.tif
Fig. 27 (A) Cycle for CO2 capture by an amine and conversion to methanol. (B) Integrated CO2 capture and conversion system. Adapted from ref. 198 and 199 with permission. Copyright 2016 and 2020 American Chemical Society.

In the case of combining capture and electrocatalysis, CO2 can be captured by an aqueous solution of inorganic hydroxides to yield corresponding bicarbonates. Even though bicarbonate is commonly used as electrolyte for conventional CO2 electrolysis, it can also serve as the carbon precursor for electrochemical reduction. So far, direct electrolysis of bicarbonate has not been reported yet, but the indirect electrochemical reduction reaction of bicarbonate solution has been achieved using a bipolar membrane (BPM) as the ion-exchange membrane in a flow cell where bicarbonate could be converted to molecular CO2 due to local acidification.200 Electrolysis of the N2-saturated 3.0 M KHCO3 solution yielded CO with a FE of 81% at 25 mA cm−2, which is comparable to the conventional gaseous CO2 electrolysis in bicarbonate solution. Similarly, amines were used to capture CO2 to form carbamate adducts.201 Electrolysis of the CO2-saturated 30% (w/w) monoethanolamine (MEA) aqueous solution led to the formation of formate with FE reaching up to 60.8% in the presence of a porous Pb electrode and cetyltrimethylammonium bromide. One challenge associated with the direct electrolysis of carbamate is the electrostatic repulsion between carbamate ion and the cathode surface. To address this issue, an alkali cation could be added into the aqueous MEA solution to change the interfacial structure near the electrode, thereby improving the electron transfer from the electrode to the carbamate and the electrochemical performance, as shown in Fig. 28A and B.202 For example, adding 2 M KCl as supporting electrolyte and by using Ag as a catalyst, CO was formed at 72% FE and a current density of 50 mA cm−2 (Fig. 28C). The amine electrolyte was recycled 10 times and could still be used for the capture and conversion reaction without obvious FECO drop, demonstrating the promising stability of the electrolyte for continuous CO2 capture and conversion (Fig. 28D).


image file: d2nr02894h-f28.tif
Fig. 28 (A) Proposed interfacial structure near the electrode surface. (B) Product distribution of MEA–CO2 conversion to H2 and CO at different applied current densities, ranging from 5 mA cm−2 to 100 mA cm−2 in a flow cell system. The error bars represent the standard deviation of three independent measurements. (C) Recycling performance of the 2 M MEA with 3 M KCl electrolyte at a constant applied current density of 10 mA cm−2 heated to 30 °C in a three-electrode configuration. Products were collected within 1 h. Adapted from ref. 202 with permission. Copyright 2021 Nature Publishing Group.

5. Concluding remarks

A sustainable carbon cycle is essential for maintaining the healthy evolution of life globally. However, human activities, especially the ever-demanding energy consumption, have led to excessive depletion of fossil fuels, and severely affected the well-established equilibrium of the carbon cycle in nature. Given the threat of excessive CO2 emission, there is now a growing demand for negative carbon technologies. Carbon capture and storage as well as direct air capture are promising technologies that could be utilized to minimize and/or reduce CO2 emissions. Various adsorbent materials have been developed for CO2 capture, including aqueous hydroxides, solid alkali carbonates, organic amines, and porous materials. To date, the investigations of direct air capture adsorbents have focused more on the use of solid-supported amine materials for improved stability and recyclability. The chemical reactions between CO2 and amines ensure significant CO2 uptake even at low CO2 partial pressures with much higher selectivity. The physical adsorption strategy using porous materials is also considered as an attractive alternative to conventional chemical adsorption approaches. Moreover, the modification of metal centres and functional groups as well as pore sizes could incorporate both chemisorption and physisorption capability within one adsorbent structure, and in turn offer better CO2 adsorption capability and selectivity. Looking into the future, practical CO2 adsorbent materials that are highly active, selective, recyclable, and cost-effective are still in demand.

The reduction of CO2 into value-added chemicals and fuels is equally important to carbon neutral and sustainable energy. Thermal catalysis of CO2 hydrogenation has been attractive because H2 can be generated from water electrolysis by renewable energy. However, this method does require high temperature and pressure for the conversion to complete. Electrochemical CO2 reduction, in comparison, can be initiated by renewable electricity under ambient conditions. To lower the activation energy barrier of CO2 and to convert CO2 to value-added chemicals, active, selective, and stable catalysts need first to be developed. Catalyst–electrolyte interfaces should also be well-engineered to eliminate all interfacial and mass transport issues during the reaction. Despite the great advances made in these areas, the development of efficient catalysts still posts some serious challenges for practical applications.

Integrated CO2 capture and conversion removes the cost of CO2 release and compression and could potentially improve the overall energy efficiency of the system. Recently, the feasibility and potential benefits of integrated CO2 capture and conversion systems have been demonstrated. But still, there is much to do in research and development to uncover the fundamental mechanisms that lead to efficient transformation of the captured CO2 to the targeted carbon products. Once the new catalysts and the reduction processes are materialized, coupling CO2 conversion with direct air capture will become a true integrated technology for realizing negative CO2 emission and energy sustainability.

Conflicts of interest

There are no conflicts of interest to declare.

Acknowledgements

Recent work was supported by the NSF under Grants CHE-2102290 (Brown), CHE-2102245 (the University of Connecticut), and by Brown's Office of Vice President for Research.

References

  1. T. J. Battin, S. Luyssaert, L. A. Kaplan, A. K. Aufdenkampe, A. Richter and L. J. Tranvik, Nat. Geosci., 2009, 2, 598–600 CrossRef CAS.
  2. R. A. Berner, Nature, 2003, 426, 323–326 CrossRef CAS PubMed.
  3. E. T. Mitchard, Nature, 2018, 559, 527–534 CrossRef CAS PubMed.
  4. E. Eriksson and P. Welander, Tellus, 1956, 8, 154–175 CrossRef.
  5. J. E. Bauer, W.-J. Cai, P. A. Raymond, T. S. Bianchi, C. S. Hopkinson and P. A. Regnier, Nature, 2013, 504, 61–70 CrossRef CAS PubMed.
  6. W. H. Schlesinger, Soils and Global Change, 1995, vol. 25, pp. 9–25 Search PubMed.
  7. P. M. Cox, R. A. Betts, C. D. Jones, S. A. Spall and I. J. Totterdell, Nature, 2000, 408, 184–187 CrossRef CAS PubMed.
  8. M. Reichstein, M. Bahn, P. Ciais, D. Frank, M. D. Mahecha, S. I. Seneviratne, J. Zscheischler, C. Beer, N. Buchmann and D. C. Frank, Nature, 2013, 500, 287–295 CrossRef CAS PubMed.
  9. J. Randers, 2052: A global forecast for the next forty years, Chelsea Green Publishing, 2012 Search PubMed.
  10. J. C. Zachos, G. R. Dickens and R. E. Zeebe, Nature, 2008, 451, 279–283 CrossRef CAS PubMed.
  11. M. Anwar, A. Fayyaz, N. Sohail, M. Khokhar, M. Baqar, W. Khan, K. Rasool, M. Rehan and A. Nizami, J. Environ. Manage., 2018, 226, 131–144 CrossRef CAS PubMed.
  12. B. Li, Y. Duan, D. Luebke and B. Morreale, Appl. Energy, 2013, 102, 1439–1447 CrossRef CAS.
  13. S. Chuhadiya, D. Suthar, S. Patel and M. Dhaka, Coord. Chem. Rev., 2021, 446, 214115 CrossRef CAS.
  14. E. S. Sanz-Perez, C. R. Murdock, S. A. Didas and C. W. Jones, Chem. Rev., 2016, 116, 11840–11876 CrossRef CAS PubMed.
  15. X. Shi, H. Xiao, H. Azarabadi, J. Song, X. Wu, X. Chen and K. S. Lackner, Angew. Chem., Int. Ed., 2020, 59, 6984–7006 CrossRef CAS PubMed.
  16. C. R. Murdock, S. Didas and C. W. Jones, Chem. Rev., 2016, 116(19), 11840–11876 CrossRef PubMed.
  17. G. Singh, J. Lee, A. Karakoti, R. Bahadur, J. Yi, D. Zhao, K. AlBahily and A. Vinu, Chem. Soc. Rev., 2020, 49, 4360–4404 RSC.
  18. S. Zeng, X. Zhang, L. Bai, X. Zhang, H. Wang, J. Wang, D. Bao, M. Li, X. Liu and S. Zhang, Chem. Rev., 2017, 117, 9625–9673 CrossRef CAS PubMed.
  19. G. Kupgan, L. J. Abbott, K. E. Hart and C. M. Colina, Chem. Rev., 2018, 118, 5488–5538 CrossRef CAS PubMed.
  20. J. K. Stolaroff, D. W. Keith and G. V. Lowry, Environ. Sci. Technol., 2008, 42, 2728–2735 CrossRef CAS PubMed.
  21. M. T. Dunstan, F. Donat, A. H. Bork, C. P. Grey and C. R. Müller, Chem. Rev., 2021, 121, 12681–12745 CrossRef CAS PubMed.
  22. X. Yang, R. J. Rees, W. Conway, G. Puxty, Q. Yang and D. A. Winkler, Chem. Rev., 2017, 117, 9524–9593 CrossRef CAS PubMed.
  23. A. Goeppert, M. Czaun, G. S. Prakash and G. A. Olah, Energy Environ. Sci., 2012, 5, 7833–7853 RSC.
  24. P. Bollini, S. A. Didas and C. W. Jones, J. Mater. Chem., 2011, 21, 15100–15120 RSC.
  25. W. Li, S. Choi, J. H. Drese, M. Hornbostel, G. Krishnan, P. M. Eisenberger and C. W. Jones, ChemSusChem, 2010, 3, 899–903 CrossRef CAS PubMed.
  26. S. A. Didas, Structural properties of aminosilica materials for CO2 capture, Georgia Institute of Technology, 2014 Search PubMed.
  27. N. R. Stuckert and R. T. Yang, Environ. Sci. Technol., 2011, 45, 10257–10264 CrossRef CAS PubMed.
  28. O. K. Farha, I. Eryazici, N. C. Jeong, B. G. Hauser, C. E. Wilmer, A. A. Sarjeant, R. Q. Snurr, S. T. Nguyen, A. O. z. r. Yazaydın and J. T. Hupp, J. Am. Chem. Soc., 2012, 134, 15016–15021 CrossRef CAS PubMed.
  29. Y. Z. Li, H. H. Wang, H. Y. Yang, L. Hou, Y. Y. Wang and Z. Zhu, Chem. – Eur. J., 2018, 24, 865–871 CrossRef CAS PubMed.
  30. A. Modak and S. Jana, Microporous Mesoporous Mater., 2019, 276, 107–132 CrossRef CAS.
  31. A. O. Yazaydın, A. I. Benin, S. A. Faheem, P. Jakubczak, J. J. Low, R. R. Willis and R. Q. Snurr, Chem. Mater., 2009, 21, 1425–1430 CrossRef.
  32. D.-A. Yang, H.-Y. Cho, J. Kim, S.-T. Yang and W.-S. Ahn, Energy Environ. Sci., 2012, 5, 6465–6473 RSC.
  33. X. Kong, E. Scott, W. Ding, J. A. Mason, J. R. Long and J. A. Reimer, J. Am. Chem. Soc., 2012, 134, 14341–14344 CrossRef CAS PubMed.
  34. R. Poloni, K. Lee, R. F. Berger, B. Smit and J. B. Neaton, J. Phys. Chem. Lett., 2014, 5, 861–865 CrossRef CAS PubMed.
  35. H. Li, K. Wang, Z. Hu, Y.-P. Chen, W. Verdegaal, D. Zhao and H.-C. Zhou, J. Mater. Chem. A, 2019, 7, 7867–7874 RSC.
  36. A. M. Fracaroli, H. Furukawa, M. Suzuki, M. Dodd, S. Okajima, F. Gándara, J. A. Reimer and O. M. Yaghi, J. Am. Chem. Soc., 2014, 136, 8863–8866 CrossRef CAS PubMed.
  37. R. W. Flaig, T. M. Osborn Popp, A. M. Fracaroli, E. A. Kapustin, M. J. Kalmutzki, R. M. Altamimi, F. Fathieh, J. A. Reimer and O. M. Yaghi, J. Am. Chem. Soc., 2017, 139, 12125–12128 CrossRef CAS PubMed.
  38. S.-J. Bao, R. Krishna, Y.-B. He, J.-S. Qin, Z.-M. Su, S.-L. Li, W. Xie, D.-Y. Du, W.-W. He and S.-R. Zhang, J. Mater. Chem. A, 2015, 3, 7361–7367 RSC.
  39. Z. Hu, Y. Wang, S. Farooq and D. Zhao, AIChE J., 2017, 63, 4103–4114 CrossRef CAS.
  40. W. Meng, Y. Zeng, Z. Liang, W. Guo, C. Zhi, Y. Wu, R. Zhong, C. Qu and R. Zou, ChemSusChem, 2018, 11, 3751–3757 CrossRef CAS PubMed.
  41. B. L. Suh, S. Lee and J. Kim, J. Phys. Chem. C, 2017, 121, 24444–24451 CrossRef CAS.
  42. S. Nakamura, M. Hatakeyama, Y. Wang, K. Ogata and K. Fujii, in Advances in CO2 Capture, Sequestration, and Conversion, ACS Publications, 2015, pp. 123–134 Search PubMed.
  43. J. M. Weber, Int. Rev. Phys. Chem., 2014, 33, 489–519 Search PubMed.
  44. L. G. Dodson, M. C. Thompson and J. M. Weber, Annu. Rev. Phys. Chem., 2018, 69, 231–252 CrossRef CAS PubMed.
  45. B. J. Knurr and J. M. Weber, J. Phys. Chem. A, 2014, 118, 4056–4062 CrossRef CAS PubMed.
  46. B. J. Knurr and J. M. Weber, J. Phys. Chem. A, 2014, 118, 8753–8757 CrossRef CAS PubMed.
  47. M. C. Thompson, L. G. Dodson and J. M. Weber, J. Phys. Chem. A, 2017, 121, 4132–4138 CrossRef CAS PubMed.
  48. B. M. Tackett, E. Gomez and J. G. Chen, Nat. Catal., 2019, 2, 381–386 CrossRef CAS.
  49. C. Song, Catal. Today, 2006, 115, 2–32 CrossRef CAS.
  50. W. Li, H. Wang, X. Jiang, J. Zhu, Z. Liu, X. Guo and C. Song, RSC Adv., 2018, 8, 7651–7669 RSC.
  51. Y. A. Daza and J. N. Kuhn, RSC Adv., 2016, 6, 49675–49691 RSC.
  52. T. Riedel, G. Schaub, K.-W. Jun and K.-W. Lee, Ind. Eng. Chem. Res., 2001, 40, 1355–1363 CrossRef CAS.
  53. M. J. L. Ginés, A. J. Marchi and C. R. Apesteguía, Appl. Catal., A, 1997, 154, 155–171 CrossRef.
  54. S.-I. Fujita, M. Usui and N. Takezawa, J. Catal., 1992, 134, 220–225 CrossRef CAS.
  55. C.-S. Chen, W.-H. Cheng and S.-S. Lin, Catal. Lett., 2000, 68, 45–48 CrossRef CAS.
  56. M. D. Porosoff, B. Yan and J. G. Chen, Energy Environ. Sci., 2016, 9, 62–73 RSC.
  57. S. Kattel, P. Liu and J. G. Chen, J. Am. Chem. Soc., 2017, 139, 9739–9754 CrossRef CAS PubMed.
  58. M. D. Porosoff and J. G. Chen, J. Catal., 2013, 301, 30–37 CrossRef CAS.
  59. J. C. Matsubu, V. N. Yang and P. Christopher, J. Am. Chem. Soc., 2015, 137, 3076–3084 CrossRef CAS PubMed.
  60. X. Zhang, X. Zhu, L. Lin, S. Yao, M. Zhang, X. Liu, X. Wang, Y.-W. Li, C. Shi and D. Ma, ACS Catal., 2017, 7, 912–918 CrossRef CAS.
  61. S.-C. Yang, S. H. Pang, T. P. Sulmonetti, W.-N. Su, J.-F. Lee, B.-J. Hwang and C. W. Jones, ACS Catal., 2018, 8, 12056–12066 CrossRef CAS.
  62. X. Jiang, X. Nie, X. Guo, C. Song and J. G. Chen, Chem. Rev., 2020, 120, 7984–8034 CrossRef CAS PubMed.
  63. G. A. Olah, A. Goeppert and G. K. S. Prakash, in Beyond Oil and Gas: The Methanol Economy, Wiley-VCH, Weinheim, 2009, pp. 179–184 Search PubMed.
  64. S.-T. Bai, G. De Smet, Y. Liao, R. Sun, C. Zhou, M. Beller, B. U. W. Maes and B. F. Sels, Chem. Soc. Rev., 2021, 50, 4259–4298 RSC.
  65. J. S. Lee, K. H. Lee, S. Y. Lee and Y. G. Kim, J. Catal., 1993, 144, 414–424 CrossRef CAS.
  66. S. Kattel, P. J. Ramírez, J. G. Chen, J. A. Rodriguez and P. Liu, Science, 2017, 355, 1296–1299 CrossRef CAS PubMed.
  67. M. Behrens, F. Studt, I. Kasatkin, S. Kühl, M. Hävecker, F. Abild-Pedersen, S. Zander, F. Girgsdies, P. Kurr, B.-L. Kniep, M. Tovar, R. W. Fischer, J. K. Nørskov and R. Schlögl, Science, 2012, 336, 893–897 CrossRef CAS PubMed.
  68. E. L. Kunkes, F. Studt, F. Abild-Pedersen, R. Schlögl and M. Behrens, J. Catal., 2015, 328, 43–48 CrossRef CAS.
  69. S. Tada, S. Kayamori, T. Honma, H. Kamei, A. Nariyuki, K. Kon, T. Toyao, K.-I. Shimizu and S. Satokawa, ACS Catal., 2018, 8, 7809–7819 CrossRef CAS.
  70. B. Rungtaweevoranit, J. Baek, J. R. Araujo, B. S. Archanjo, K. M. Choi, O. M. Yaghi and G. A. Somorjai, Nano Lett., 2016, 16, 7645–7649 CrossRef CAS PubMed.
  71. F. Studt, I. Sharafutdinov, F. Abild-Pedersen, C. F. Elkjær, J. S. Hummelshøj, S. Dahl, I. Chorkendorff and J. K. Nørskov, Nat. Chem., 2014, 6, 320–324 CrossRef CAS PubMed.
  72. A. García-Trenco, A. Regoutz, E. R. White, D. J. Payne, M. S. P. Shaffer and C. K. Williams, Appl. Catal., B, 2018, 220, 9–18 CrossRef.
  73. O. Martin, A. J. Martín, C. Mondelli, S. Mitchell, T. F. Segawa, R. Hauert, C. Drouilly, D. Curulla-Ferré and J. Pérez-Ramírez, Angew. Chem., Int. Ed., 2016, 55, 6261–6265 CrossRef CAS PubMed.
  74. E. C. Ra, K. Y. Kim, E. H. Kim, H. Lee, K. An and J. S. Lee, ACS Catal., 2020, 10, 11318–11345 CrossRef CAS.
  75. I. Amghizar, L. A. Vandewalle, K. M. Van Geem and G. B. Marin, Engineering, 2017, 3, 171–178 CrossRef CAS.
  76. W. Zhou, K. Cheng, J. Kang, C. Zhou, V. Subramanian, Q. Zhang and Y. Wang, Chem. Soc. Rev., 2019, 48, 3193–3228 RSC.
  77. R.-P. Ye, J. Ding, W. Gong, M. D. Argyle, Q. Zhong, Y. Wang, C. K. Russell, Z. Xu, A. G. Russell, Q. Li, M. Fan and Y.-G. Yao, Nat. Commun., 2019, 10, 5698 CrossRef CAS PubMed.
  78. J. Chai, R. Pestman, W. Chen, N. Donkervoet, A. I. Dugulan, Z. Men, P. Wang and E. J. Hensen, ACS Catal., 2022, 12, 2877–2887 CrossRef CAS.
  79. Q.-Y. Liu, C. Shang and Z.-P. Liu, J. Phys. Chem. Lett., 2022, 13, 3342–3352 CrossRef CAS PubMed.
  80. A. E. Rashed, A. Nasser, M. F. Elkady, Y. Matsushita and A. A. El-Moneim, ACS Omega, 2022, 7, 8403–8419 CrossRef CAS PubMed.
  81. R. W. Dorner, D. R. Hardy, F. W. Williams and H. D. Willauer, Energy Environ. Sci., 2010, 3, 884–890 RSC.
  82. J. Wei, J. Sun, Z. Wen, C. Fang, Q. Ge and H. Xu, Catal. Sci. Technol., 2016, 6, 4786–4793 RSC.
  83. A. Ramirez, L. Gevers, A. Bavykina, S. Ould-Chikh and J. Gascon, ACS Catal., 2018, 8, 9174–9182 CrossRef CAS.
  84. W. Wang, X. Jiang, X. Wang and C. Song, Ind. Eng. Chem. Res., 2018, 57, 4535–4542 CrossRef CAS.
  85. R. Satthawong, N. Koizumi, C. Song and P. Prasassarakich, J. CO2 Util., 2013, 3–4, 102–106 CrossRef CAS.
  86. J. Wang, Z. You, Q. Zhang, W. Deng and Y. Wang, Catal. Today, 2013, 215, 186–193 CrossRef CAS.
  87. S. Gupta, V. K. Jain and D. Jagadeesan, ChemNanoMat, 2016, 2, 989–996 CrossRef CAS.
  88. D. Wang, Z. Xie, M. D. Porosoff and J. G. Chen, Chem, 2021, 7, 2277–2311 CAS.
  89. K. B. Tan, G. Zhan, D. Sun, J. Huang and Q. Li, J. Mater. Chem. A, 2021, 9, 5197–5231 RSC.
  90. L. Fan, C. Xia, F. Yang, J. Wang, H. Wang and Y. Lu, Sci. Adv., 2020, 6, eaay3111 CrossRef CAS PubMed.
  91. Z. Sun, Y. Hu, D. Zhou, M. Sun, S. Wang and W. Chen, ACS Energy Lett., 2021, 6, 3992–4022 CrossRef CAS.
  92. T. K. Todorova, M. W. Schreiber and M. Fontecave, ACS Catal., 2019, 10, 1754–1768 CrossRef.
  93. Z. Yin, G. T. R. Palmore and S. Sun, Trends Chem., 2019, 1, 739–750 CrossRef CAS.
  94. D. Johnson, Z. Qiao and A. Djire, ACS Appl. Energy Mater., 2021, 4, 8661–8684 CrossRef CAS.
  95. S. Nitopi, E. Bertheussen, S. B. Scott, X. Liu, A. K. Engstfeld, S. Horch, B. Seger, I. E. Stephens, K. Chan and C. Hahn, Chem. Rev., 2019, 119, 7610–7672 CrossRef CAS PubMed.
  96. H. Zhang, Y. Zhang, Y. Li, S. Ahn, G. T. R. Palmore, J. Fu, A. A. Peterson and S. Sun, Nanoscale, 2019, 11, 12075–12079 RSC.
  97. J. Fu, W. Zhu, Y. Chen, Z. Yin, Y. Li, J. Liu, H. Zhang, J. J. Zhu and S. Sun, Angew. Chem., 2019, 131, 14238–14241 CrossRef.
  98. W. Zhu, R. Michalsky, O. N. Metin, H. Lv, S. Guo, C. J. Wright, X. Sun, A. A. Peterson and S. Sun, J. Am. Chem. Soc., 2013, 135, 16833–16836 CrossRef CAS PubMed.
  99. W. Zhu, Y.-J. Zhang, H. Zhang, H. Lv, Q. Li, R. Michalsky, A. A. Peterson and S. Sun, J. Am. Chem. Soc., 2014, 136, 16132–16135 CrossRef CAS PubMed.
  100. D. Gao, H. Zhou, J. Wang, S. Miao, F. Yang, G. Wang, J. Wang and X. Bao, J. Am. Chem. Soc., 2015, 137, 4288–4291 CrossRef CAS PubMed.
  101. C. Kim, H. S. Jeon, T. Eom, M. S. Jee, H. Kim, C. M. Friend, B. K. Min and Y. J. Hwang, J. Am. Chem. Soc., 2015, 137, 13844–13850 CrossRef CAS PubMed.
  102. Q. Li, J. Fu, W. Zhu, Z. Chen, B. Shen, L. Wu, Z. Xi, T. Wang, G. Lu and J.-J. Zhu, J. Am. Chem. Soc., 2017, 139, 4290–4293 CrossRef CAS PubMed.
  103. P. Hou, X. Wang, Z. Wang and P. Kang, ACS Appl. Mater. Interfaces, 2018, 10, 38024–38031 CrossRef CAS PubMed.
  104. S. Zhang, P. Kang and T. J. Meyer, J. Am. Chem. Soc., 2014, 136, 1734–1737 CrossRef CAS PubMed.
  105. D. Gao, H. Zhou, F. Cai, J. Wang, G. Wang and X. Bao, ACS Catal., 2018, 8, 1510–1519 CrossRef CAS.
  106. C. W. Lee, J. S. Hong, K. D. Yang, K. Jin, J. H. Lee, H.-Y. Ahn, H. Seo, N.-E. Sung and K. T. Nam, ACS Catal., 2018, 8, 931–937 CrossRef CAS.
  107. X. Sun, L. Lu, Q. Zhu, C. Wu, D. Yang, C. Chen and B. Han, Angew. Chem., Int. Ed., 2018, 57, 2427–2431 CrossRef CAS PubMed.
  108. Q. Zhu, X. Sun, D. Yang, J. Ma, X. Kang, L. Zheng, J. Zhang, Z. Wu and B. Han, Nat. Commun., 2019, 10, 1–11 CrossRef PubMed.
  109. C. W. Li and M. W. Kanan, J. Am. Chem. Soc., 2012, 134, 7231–7234 CrossRef CAS PubMed.
  110. Y. Zhao, X. Zu, R. Chen, X. Li, Y. Jiang, Z. Wang, S. Wang, Y. Wu, Y. Sun and Y. Xie, J. Am. Chem. Soc., 2022, 144, 10446–10454 CrossRef CAS PubMed.
  111. S. Mu, H. Lu, Q. Wu, L. Li, R. Zhao, C. Long and C. Cui, Nat. Commun., 2022, 13, 1–8 Search PubMed.
  112. X. Wang, K. Klingan, M. Klingenhof, T. Möller, J. Ferreira de Araújo, I. Martens, A. Bagger, S. Jiang, J. Rossmeisl and H. Dau, Nat. Commun., 2021, 12, 1–12 CrossRef PubMed.
  113. S. H. Lee, J. C. Lin, M. Farmand, A. T. Landers, J. T. Feaster, J. E. Avilés Acosta, J. W. Beeman, Y. Ye, J. Yano and A. Mehta, J. Am. Chem. Soc., 2020, 143, 588–592 CrossRef PubMed.
  114. S.-C. Lin, C.-C. Chang, S.-Y. Chiu, H.-T. Pai, T.-Y. Liao, C.-S. Hsu, W.-H. Chiang, M.-K. Tsai and H. M. Chen, Nat. Commun., 2020, 11, 1–12 CrossRef PubMed.
  115. Z. Yin, C. Yu, Z. Zhao, X. Guo, M. Shen, N. Li, M. Muzzio, J. Li, H. Liu and H. Lin, Nano Lett., 2019, 19, 8658–8663 CrossRef CAS PubMed.
  116. S. Meshitsuka, M. Ichikawa and K. Tamaru, J. Chem. Soc., Chem. Commun., 1974, 158–159 RSC.
  117. M. Li, H. Wang, W. Luo, P. C. Sherrell, J. Chen and J. Yang, Adv. Mater., 2020, 32, 2001848 CrossRef CAS PubMed.
  118. J. Yang, Z. Qiu, C. Zhao, W. Wei, W. Chen, Z. Li, Y. Qu, J. Dong, J. Luo and Z. Li, Angew. Chem., Int. Ed., 2018, 57, 14095–14100 CrossRef CAS PubMed.
  119. X. Wang, Z. Chen, X. Zhao, T. Yao, W. Chen, R. You, C. Zhao, G. Wu, J. Wang and W. Huang, Angew. Chem., 2018, 130, 1962–1966 CrossRef.
  120. Q. Qu, S. Ji, Y. Chen, D. Wang and Y. Li, Chem. Sci., 2021, 12, 4201–4215 RSC.
  121. Y. Wang, H. Su, Y. He, L. Li, S. Zhu, H. Shen, P. Xie, X. Fu, G. Zhou and C. Feng, Chem. Rev., 2020, 120, 12217–12314 CrossRef CAS PubMed.
  122. C. Shi, H. A. Hansen, A. C. Lausche and J. K. Nørskov, Phys. Chem. Chem. Phys., 2014, 16, 4720–4727 RSC.
  123. Y. Hori, H. Wakebe, T. Tsukamoto and O. Koga, Electrochim. Acta, 1994, 39, 1833–1839 CrossRef CAS.
  124. A. Bagger, W. Ju, A. S. Varela, P. Strasser and J. Rossmeisl, ChemPhysChem, 2017, 18, 3266–3273 CrossRef CAS PubMed.
  125. C. W. Lee, K. D. Yang, D. H. Nam, J. H. Jang, N. H. Cho, S. W. Im and K. T. Nam, Adv. Mater., 2018, 30, 1704717 CrossRef PubMed.
  126. H. Yano, F. Shirai, M. Nakayama and K. Ogura, J. Electroanal. Chem., 2002, 533, 113–118 CrossRef CAS.
  127. T. Burdyny, P. J. Graham, Y. Pang, C.-T. Dinh, M. Liu, E. H. Sargent and D. Sinton, ACS Sustainable Chem. Eng., 2017, 5, 4031–4040 CrossRef CAS.
  128. K. Junge Puring, D. Siegmund, J. Timm, F. Möllenbruck, S. Schemme, R. Marschall and U. P. Apfel, Adv. Sustainable Syst., 2021, 5, 2000088 CrossRef CAS.
  129. Z. Zhan and L. Zhao, J. Power Sources, 2010, 195, 7250–7254 CrossRef CAS.
  130. M. Moura de Salles Pupo and R. Kortlever, ChemPhysChem, 2019, 20, 2926–2935 CrossRef CAS PubMed.
  131. J.-B. Vennekötter, T. Scheuermann, R. Sengpiel and M. Wessling, J. CO2 Util., 2019, 32, 202–213 CrossRef.
  132. S. Lu, Y. Wang, H. Xiang, H. Lei, B. B. Xu, L. Xing, E. H. Yu and T. X. Liu, J. Energy Storage, 2022, 52, 104764 CrossRef.
  133. D. T. Hofsommer, Y. Liang, S. S. Uttarwar, M. Gautam, S. Pishgar, S. Gulati, C. A. Grapperhaus and J. M. Spurgeon, ChemSusChem, 2022, 15, e202102289 CrossRef CAS PubMed.
  134. S. Zhang, P. Kang, S. Ubnoske, M. K. Brennaman, N. Song, R. L. House, J. T. Glass and T. J. Meyer, J. Am. Chem. Soc., 2014, 136, 7845–7848 CrossRef CAS PubMed.
  135. M. S. Xie, B. Y. Xia, Y. Li, Y. Yan, Y. Yang, Q. Sun, S. H. Chan, A. Fisher and X. Wang, Energy Environ. Sci., 2016, 9, 1687–1695 RSC.
  136. Y. Qiu, H. Zhong, W. Xu, T. Zhang, X. Li and H. Zhang, J. Mater. Chem. A, 2019, 7, 5453–5462 RSC.
  137. L. Jin, B. Liu, P. Wang, H. Yao, L. A. Achola, P. Kerns, A. Lopes, Y. Yang, J. Ho and A. Moewes, Nanoscale, 2018, 10, 14678–14686 RSC.
  138. L. Jin, E. Shaaban, S. Bamonte, D. Cintron, S. Shuster, L. Zhang, G. Li and J. He, ACS Appl. Mater. Interfaces, 2021, 13, 38595–38603 CrossRef CAS PubMed.
  139. X. Su, Y. Sun, L. Jin, L. Zhang, Y. Yang, P. Kerns, B. Liu, S. Li and J. He, Appl. Catal., B, 2020, 269, 118800 CrossRef CAS.
  140. J. C. Bui, C. Kim, A. J. King, O. Romiluyi, A. Kusoglu, A. Z. Weber and A. T. Bell, Acc. Chem. Res., 2022, 55, 484–494 CrossRef CAS PubMed.
  141. Q. Fan, M. Zhang, M. Jia, S. Liu, J. Qiu and Z. Sun, Mater. Today Energy, 2018, 10, 280–301 CrossRef.
  142. L. Fan, C.-Y. Liu, P. Zhu, C. Xia, X. Zhang, Z.-Y. Wu, Y. Lu, T. P. Senftle and H. Wang, Joule, 2022, 6, 205–220 CrossRef CAS.
  143. Y. J. Sa, C. W. Lee, S. Y. Lee, J. Na, U. Lee and Y. J. Hwang, Chem. Soc. Rev., 2020, 49, 6632–6665 RSC.
  144. H. Hashiba, L.-C. Weng, Y. Chen, H. K. Sato, S. Yotsuhashi, C. Xiang and A. Z. Weber, J. Phys. Chem. C, 2018, 122, 3719–3726 CrossRef CAS.
  145. R. Casebolt, K. W. Kimura, K. Levine, J. A. Cimada DaSilva, J. Kim, T. A. Dunbar, J. Suntivich and T. Hanrath, ChemElectroChem, 2021, 8, 681–688 CrossRef CAS.
  146. A. S. Varela, M. Kroschel, T. Reier and P. Strasser, Catal. Today, 2016, 260, 8–13 CrossRef CAS.
  147. M. R. Singh, E. L. Clark and A. T. Bell, Phys. Chem. Chem. Phys., 2015, 17, 18924–18936 RSC.
  148. C. Kim, J. C. Bui, X. Luo, J. K. Cooper, A. Kusoglu, A. Z. Weber and A. T. Bell, Nat. Energy, 2021, 6, 1026–1034 CrossRef CAS.
  149. M. T. Koper, Chem. Sci., 2013, 4, 2710–2723 RSC.
  150. X. Lu, C. Zhu, Z. Wu, J. Xuan, J. S. Francisco and H. Wang, J. Am. Chem. Soc., 2020, 142, 15438–15444 CrossRef CAS PubMed.
  151. M. König, J. Vaes, E. Klemm and D. Pant, iScience, 2019, 19, 135–160 CrossRef PubMed.
  152. J. Li, Y. Kuang, Y. Meng, X. Tian, W.-H. Hung, X. Zhang, A. Li, M. Xu, W. Zhou, C.-S. Ku, C.-Y. Chiang, G. Zhu, J. Guo, X. Sun and H. Dai, J. Am. Chem. Soc., 2020, 142, 7276–7282 CrossRef CAS PubMed.
  153. A. S. Hall, Y. Yoon, A. Wuttig and Y. Surendranath, J. Am. Chem. Soc., 2015, 137, 14834–14837 CrossRef CAS PubMed.
  154. Y. Zou and S. Wang, Adv. Sci., 2021, 8, 2003579 CrossRef CAS PubMed.
  155. R. Kas, R. Kortlever, H. Yılmaz, M. T. M. Koper and G. Mul, ChemElectroChem, 2015, 2, 354–358 CrossRef CAS.
  156. M. R. Singh, Y. Kwon, Y. Lum, J. W. Ager and A. T. Bell, J. Am. Chem. Soc., 2016, 138, 13006–13012 CrossRef CAS PubMed.
  157. J. Resasco, L. D. Chen, E. Clark, C. Tsai, C. Hahn, T. F. Jaramillo, K. Chan and A. T. Bell, J. Am. Chem. Soc., 2017, 139, 11277–11287 CrossRef CAS PubMed.
  158. D. Gao, F. Scholten and B. Roldan Cuenya, ACS Catal., 2017, 7, 5112–5120 CrossRef CAS.
  159. J. Resasco, Y. Lum, E. Clark, J. Z. Zeledon and A. T. Bell, ChemElectroChem, 2018, 5, 1064–1072 CrossRef CAS.
  160. Y. Ma, J. Wang, J. Yu, J. Zhou, X. Zhou, H. Li, Z. He, H. Long, Y. Wang and P. Lu, Matter, 2021, 4, 888–926 CrossRef CAS.
  161. F. Li, A. Thevenon, A. Rosas-Hernández, Z. Wang, Y. Li, C. M. Gabardo, A. Ozden, C. T. Dinh, J. Li and Y. Wang, Nature, 2020, 577, 509–513 CrossRef CAS PubMed.
  162. L. Zhang, Z. Wei, S. Thanneeru, M. Meng, M. Kruzyk, G. Ung, B. Liu and J. He, Angew. Chem., 2019, 131, 15981–15987 CrossRef.
  163. Y. Wu, Z. Jiang, X. Lu, Y. Liang and H. Wang, Nature, 2019, 575, 639–642 CrossRef CAS PubMed.
  164. Z. Cao, D. Kim, D. Hong, Y. Yu, J. Xu, S. Lin, X. Wen, E. M. Nichols, K. Jeong, J. A. Reimer, P. Yang and C. J. Chang, J. Am. Chem. Soc., 2016, 138, 8120–8125 CrossRef CAS PubMed.
  165. K. Jiang, Y. Huang, G. Zeng, F. M. Toma, W. A. Goddard III and A. T. Bell, ACS Energy Lett., 2020, 5, 1206–1214 CrossRef CAS.
  166. H. Mistry, Y. W. Choi, A. Bagger, F. Scholten, C. S. Bonifacio, I. Sinev, N. J. Divins, I. Zegkinoglou, H. S. Jeon and K. Kisslinger, Angew. Chem., 2017, 129, 11552–11556 CrossRef.
  167. Q. Zhu, C. J. Murphy and L. R. Baker, J. Am. Chem. Soc., 2022, 144, 2829–2840 CrossRef CAS PubMed.
  168. D. Wakerley, S. Lamaison, F. Ozanam, N. Menguy, D. Mercier, P. Marcus, M. Fontecave and V. Mougel, Nat. Mater., 2019, 18, 1222–1227 CrossRef CAS PubMed.
  169. D. Kim, S. Yu, F. Zheng, I. Roh, Y. Li, S. Louisia, Z. Qi, G. A. Somorjai, H. Frei and L.-W. Wang, Nat. Energy, 2020, 5, 1032–1042 CrossRef CAS.
  170. S. Yu, S. Louisia and P. Yang, JACS Au, 2022, 2, 562–572 CrossRef CAS PubMed.
  171. J. Tamura, A. Ono, Y. Sugano, C. Huang, H. Nishizawa and S. Mikoshiba, Phys. Chem. Chem. Phys., 2015, 17, 26072–26078 RSC.
  172. Y. Wang, H. Su, C. Xu, G. Li, L. Gell, S. Lin, Z. Tang, H. Häkkinen and N. Zheng, J. Am. Chem. Soc., 2015, 137, 4324–4327 CrossRef CAS PubMed.
  173. H. Shang, S. K. Wallentine, D. M. Hofmann, Q. Zhu, C. J. Murphy and L. R. Baker, Chem. Sci., 2020, 11, 12298–12306 RSC.
  174. F. Pan and Y. Yang, Energy Environ. Sci., 2020, 13, 2275–2309 RSC.
  175. Y. Fang and J. C. Flake, J. Am. Chem. Soc., 2017, 139, 3399–3405 CrossRef CAS PubMed.
  176. E. Barton Cole, P. S. Lakkaraju, D. M. Rampulla, A. J. Morris, E. Abelev and A. B. Bocarsly, J. Am. Chem. Soc., 2010, 132, 11539–11551 CrossRef CAS PubMed.
  177. P. De la Presa, M. Multigner, J. De la Venta, M. García and M. Ruiz-González, J. Appl. Phys., 2006, 100, 123915 CrossRef.
  178. S. Sun and H. Zeng, J. Am. Chem. Soc., 2002, 124, 8204–8205 CrossRef CAS PubMed.
  179. R. Hao, R. Xing, Z. Xu, Y. Hou, S. Gao and S. Sun, Adv. Mater., 2010, 22, 2729–2742 CrossRef CAS PubMed.
  180. V. Mazumder and S. Sun, J. Am. Chem. Soc., 2009, 131, 4588–4589 CrossRef CAS PubMed.
  181. C. Wang, H. Daimon, T. Onodera, T. Koda and S. Sun, Angew. Chem., 2008, 120, 3644–3647 CrossRef.
  182. Y. Zhao, C. Wang, Y. Liu, D. R. MacFarlane and G. G. Wallace, Adv. Energy Mater., 2018, 8, 1801400 CrossRef.
  183. C. Kim, T. Eom, M. S. Jee, H. Jung, H. Kim, B. K. Min and Y. J. Hwang, ACS Catal., 2017, 7, 779–785 CrossRef CAS.
  184. M. R. Narouz, K. M. Osten, P. J. Unsworth, R. W. Y. Man, K. Salorinne, S. Takano, R. Tomihara, S. Kaappa, S. Malola, C.-T. Dinh, J. D. Padmos, K. Ayoo, P. J. Garrett, M. Nambo, J. H. Horton, E. H. Sargent, H. Häkkinen, T. Tsukuda and C. M. Crudden, Nat. Chem., 2019, 11, 419–425 CrossRef CAS PubMed.
  185. L. Zhang, Z. Wei, M. Meng, G. Ung and J. He, J. Mater. Chem. A, 2020, 8, 15900–15908 RSC.
  186. M. N. Hopkinson, C. Richter, M. Schedler and F. Glorius, Nature, 2014, 510, 485–496 CrossRef CAS PubMed.
  187. T. Dröge and F. Glorius, Angew. Chem., Int. Ed., 2010, 49, 6940–6952 CrossRef PubMed.
  188. D. J. Nelson and S. P. Nolan, Chem. Soc. Rev., 2013, 42, 6723–6753 RSC.
  189. S. Engel, E.-C. Fritz and B. J. Ravoo, Chem. Soc. Rev., 2017, 46, 2057–2075 RSC.
  190. C. M. Crudden, J. H. Horton, M. R. Narouz, Z. Li, C. A. Smith, K. Munro, C. J. Baddeley, C. R. Larrea, B. Drevniok, B. Thanabalasingam, A. B. McLean, O. V. Zenkina, I. I. Ebralidze, Z. She, H.-B. Kraatz, N. J. Mosey, L. N. Saunders and A. Yagi, Nat. Commun., 2016, 7, 12654 CrossRef CAS PubMed.
  191. Z. Cao, J. S. Derrick, J. Xu, R. Gao, M. Gong, E. M. Nichols, P. T. Smith, X. Liu, X. Wen, C. Copéret and C. J. Chang, Angew. Chem., Int. Ed., 2018, 57, 4981–4985 CrossRef CAS PubMed.
  192. S. Sun, Nat. Energy, 2020, 5, 943–944 CrossRef.
  193. C. M. Jens, L. Müller, K. Leonhard and A. Bardow, ACS Sustainable Chem. Eng., 2019, 7, 12270–12280 CAS.
  194. S. Kar, A. Goeppert and G. K. S. Prakash, Acc. Chem. Res., 2019, 52, 2892–2903 CrossRef CAS PubMed.
  195. I. Sullivan, A. Goryachev, I. A. Digdaya, X. Li, H. A. Atwater, D. A. Vermaas and C. Xiang, Nat. Catal., 2021, 4, 952–958 CrossRef CAS.
  196. Y.-N. Li, L.-N. He, A.-H. Liu, X.-D. Lang, Z.-Z. Yang, B. Yu and C.-R. Luan, Green Chem., 2013, 15, 2825–2829 RSC.
  197. J. Kothandaraman, A. Goeppert, M. Czaun, G. A. Olah and G. K. Surya Prakash, Green Chem., 2016, 18, 5831–5838 RSC.
  198. J. Kothandaraman, A. Goeppert, M. Czaun, G. A. Olah and G. K. S. Prakash, J. Am. Chem. Soc., 2016, 138, 778–781 CrossRef CAS PubMed.
  199. R. Sen, A. Goeppert, S. Kar and G. K. S. Prakash, J. Am. Chem. Soc., 2020, 142, 4544–4549 CrossRef CAS PubMed.
  200. T. Li, E. W. Lees, M. Goldman, D. A. Salvatore, D. M. Weekes and C. P. Berlinguette, Joule, 2019, 3, 1487–1497 CrossRef CAS.
  201. L. Chen, F. Li, Y. Zhang, C. L. Bentley, M. Horne, A. M. Bond and J. Zhang, ChemSusChem, 2017, 10, 4109–4118 CrossRef CAS PubMed.
  202. G. Lee, Y. C. Li, J.-Y. Kim, T. Peng, D.-H. Nam, A. Sedighian Rasouli, F. Li, M. Luo, A. H. Ip, Y.-C. Joo and E. H. Sargent, Nat. Energy, 2021, 6, 46–53 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2022